You are on page 1of 103

1

Technical report

Assessment and reporting on


soil erosion
Background and workshop report

Prepared by:
Anne Gobin, Gerard Govers, Katholieke Universiteit Leuven
Robert Jones, Joint Research Centre
Mike Kirkby, University of Leeds
Costas Kosmas, Agricultural University of Athens

Project Manager:
Anna Rita Gentile
European Environment Agency

94

Assessment and reporting on soil erosion

Cover design: Rolf Kuchling, EEA


Layout: Brandenborg a/s

Legal notice
The contents of this report do not necessarily reflect the official opinion of the European Commission
or other European Communities institutions. Neither the European Environment Agency nor any
person or company acting on behalf of the Agency is responsible for the use that may be made of
the information contained in this report.
A great deal of additional information on the European Union is available on the Internet.
It can be accessed through the Europa server (http://europa.eu.int)
EEA, Copenhagen, 2003
Reproduction is authorised provided the source is acknowledged.
ISBN: 92-9167-519-9

Environmental production:
This publication is printed according to the highest environmental standards.
Printed by Scanprint a/s:
Environment Certificate: ISO 14001
Quality Certificate: ISO 9001:2000
EMAS registered; licence no. DK-S-000015
Approved for printing with the Nordic Swan environmental label, licence no. 541 055
Paper:
100 % recycled and chlorine-free bleached paper
The Nordic Swan Label
Printed in Denmark

European Environment Agency


Kongens Nytorv 6
DK-1050 Copenhagen K
Tel. (45) 33 36 71 00
Fax (45) 33 36 71 99
E-mail: eea@eea.eu.int
Internet: http://www.eea.eu.int

Contents

Contents
Executive summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

10

1.1. Scope of the report . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

10

1.2. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

10

1.3. Policy developments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

11

1.4. Objectives and methodology of the review . . . . . . . . . . . . . . . . . . . . . .

11

1.5. Soil erosion in Europe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

12

Part I Assessment and reporting on soil erosion . . . . . . . . . . . . . . . . . . . . . . . .

14

2. A European framework for the assessment and monitoring of soil . . . . . . . . .

14

2.1. The assessment framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

14

2.2. The DPSIR assessment framework applied to soil erosion . . . . . . . . . . .

17

2.3. Is the proposed DPSIR assessment framework adequate to comprehend


soil erosion? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

17

2.4. EEA typology of indicators applied to soil erosion . . . . . . . . . . . . . . . . .

19

Indicators of soil erosion and data availability . . . . . . . . . . . . . . . . . . . . . . . . .

19

2.5. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

19

2.6. Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

19

2.6.1. Indicators of driving forces and pressures . . . . . . . . . . . . . . . . . . . . .

22

2.6.2. Indicators of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

23

2.6.3. Indicators of impact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

25

2.6.4. Indicators of response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

25

2.7. Options for the future: determining the risk of soil erosion . . . . . . . . . .

25

2.7.1. Expert-based methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

26

2.7.2. Model-based methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

26

2.8. General conclusions of review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

27

3. Driving force, pressure and state indicators related to land use . . . . . . . . . . .

28

3.1. Soil erosion indicators and land use . . . . . . . . . . . . . . . . . . . . . . . . . . . .

28

3.2. Review of the proposed indicators in relation to land use intensity . . . .

29

3.3. Options for the future on relating land use and land use intensity to
soil erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

29

3.3.1. Climate characteristics affecting vegetation . . . . . . . . . . . . . . . . . . . .

29

3.3.2. Vegetation characteristics affecting soil erosion . . . . . . . . . . . . . . . . .

30

3.3.3. Management quality and human-induced factors . . . . . . . . . . . . . . .

31

3.4. Conclusions of review of indicators in relation to land use . . . . . . . . . . .

33

4. Regional assessment of the extent of soil erosion by water . . . . . . . . . . . . . .

35

4.1. Alternative assessment methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

35

4.1.1. Distributed point data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

35

4.1.2. Factor or indicator mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

36

4.1.3. Process modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

36

Assessment and reporting on soil erosion

4.2. The Corine approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

36

4.2.1. Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

37

4.2.2. Advantages and limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39

4.3. The hot-spot approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39

4.3.1. Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39

4.3.2. Advantages and limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

40

4.4. The RIVM approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

41

4.4.1. Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

42

4.4.2. Advantages and limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

42

4.5. The Glasod approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

43

4.5.1. Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

44

4.5.2. Advantages and limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

45

4.6. Comparative assessment of the four methodologies . . . . . . . . . . . . . . .

45

4.7. Options for the future . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

45

4.8. Conclusions and recommendations on implementation of regional


assessments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

46

Part II Workshop conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

48

5. Soil erosion indicators and assessment framework . . . . . . . . . . . . . . . . . . . . .

49

5.1. Operational framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

49

5.2. Soil erosion indicator work at ETC/Soil . . . . . . . . . . . . . . . . . . . . . . . . . .

49

5.3. GISCO databases and tools to derive pressure indicators for soil erosion

49

5.4. Discussion on questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

50

6. Regional and spatial assessment methods of soil erosion and data availability

51

6.1. The Glasod map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

51

6.2. The hot-spot map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

51

6.3. Regional assessment of the extent of soil erosion by water . . . . . . . . . .

51

6.4. General discussion on regional/spatial soil erosion indicators . . . . . . . .

52

7. General discussion on indicators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

54

7.1. Data availability for soil erosion indicators . . . . . . . . . . . . . . . . . . . . . . .

54

7.2. Indicators of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

54

7.3. Indicators of impact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

54

Part III Recommendations for further work . . . . . . . . . . . . . . . . . . . . . . . . . . . .

55

8. Recommendations to the EEA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

55

8.1. General recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

55

8.2. Recommendations related to the DPSIR assessment framework . . . . . .

55

8.3. General recommendations related to the proposed indicators . . . . . . .

56

8.4. Recommendations related to land use and soil erosion indicators . . . . .

57

8.5. Recommendations related to regional erosion assessment


(indicators of state) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

58

9. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

60

10. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

61

Contents

Annexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

65

Annex I List of participants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

65

Annex II Agenda . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

66

Annex III Background papers presented at the workshop . . . . . . . . . . . . . . . .

68

State of play of EEA work on soil erosion indicators . . . . . . . . . . . . . . . . . . . .

68

Soil erosion hot-spot map for Europe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Data quality issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The design and use of this map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Interpretation of the map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

71
71
72
73
73
74

Qualitative small-scale soil degradation assessment databases


The Glasod map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The Glasod map (1990) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Follow-up of Glasod / derived initiatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Methodological details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Results of the assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

74
74
76
77
79

Indicators of soil erosion at the ETC/Soil . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The indicator concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
DPSIR applied to soil erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

79
79
80
80
82

GISCO databases and tools to derive driving force/pressure indicators for


soil erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Introduction to GISCO databases and tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Overview of driving force/pressure indicators proposed by the EEA . . . . . . . . . . .
GISCO and driving force/pressure indicators . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Proposed indicator framework model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Remarks and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

83
83
85
85
85
86

Regional assessment of the impact of soil erosion by water . . . . . . . . . . . . . .


Soil erosion indicators of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The revised DPSIR assessment framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Processes of soil erosion by water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Regional assessment methods of soil erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Process modelling to assess regional soil erosion: Pesera . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

86
86
87
87
87
89
91

Data availability for soil erosion indicators at European level . . . . . . . . . . . . .


Determining the causes of soil erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Modelling soil erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Soil erosion risk assessments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Environmental indicators for soil erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

92
92
92
93
95

Annex IV Soil erosion glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

97

Annex V Processes of soil erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


99
Soil erosion by water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

Assessment and reporting on soil erosion

Abbreviations
CAP
Common agricultural policy
Corine

Coordination of information on the environment

DPSIR

Driving forces Pressures State Impact Responses

DSR

Driving forces State Responses

EEA

European Environment Agency

EFMA

European Fertiliser Manufacturers Association

EIONET

European Environmental Information and Observation Network

ETC/S

European Topic Centre on Soil

ETC/TE

European Topic Centre on Terrestrial Environment

Glasod

Global assessment of human-induced soil degradation

NDVI

Normalised difference vegetation index

OECD

Organisation for Economic Cooperation and Development

Pesera

Pan-European soil erosion risk assessment

RUSLE

Revised universal soil loss equation

UN

United Nations

UNCED

United Nations Conference on Environment and Development (Rio, 1992)

USLE

Universal soil loss equation

Acknowledgements
Special thanks to the national experts who participated in the EEA technical workshop on
indicators for soil erosion held in Copenhagen in March 2001; to Paul Campling at the
Katholieke Universiteit Leuven for his help in the organisation of the workshop; and to
Robert Evans, University of East Anglia, and Jaume Fons, Autonomous University of
Barcelona, for their useful comments.

Executive summary

Executive summary
This report has been prepared by the
Katholieke Universiteit Leuven under
contract to the European Environment
Agency (EEA) and is the final result of a
working group on indicators for soil erosion.
The working group was established by the
EEA in order to progress with the work on
soil in the interim period before the new
European Topic Centre on Terrestrial
Environment (ETC/TE) started in July 2001.
In 2001 the EEA carried out a peer review of
its work on soil, with particular reference to
the development of policy-relevant indicators
and the identification of probable problem
areas for soil degradation (hot spots) (1).
The review was in particular focused on work
on indicators for soil erosion and soil sealing,
and two associated technical workshops were
held in March 2001 to facilitate this review.
This report provides the background on and
analyses the work done by the EEA on soil
erosion in the period to 2001 and
summarises the conclusions of the workshop
on indicators for soil erosion, held in
Copenhagen on 2728 March 2001.
The purpose of the workshop was to identify
a set of recommendations concerning
reporting on soil erosion (as part of the wider
theme of soil degradation) that could then
be considered for inclusion in the work
programme for the new ETC on Terrestrial
Environment.
Soil erosion is a natural process, occurring
over geological time. Most concerns about
erosion are related to accelerated erosion,
where the natural rate has been significantly
increased by human activities such as
changes in land cover and management. This
report focuses on accelerated erosion caused
by water.
Runoff is the most important direct pressure
of severe soil erosion. Processes that
influence runoff must therefore play an
important role in any analysis of soil erosion
intensity, and measures that reduce runoff
are critical to effective soil conservation.
(1)

In Europe, soil erosion is caused mainly by


water and, to a lesser extent, by wind. In the
Mediterranean region, water erosion results
from intense seasonal rainfall on often fragile
soils located on steep slopes. The area
affected by erosion in northern Europe is
more restricted and moderate rates of water
erosion result from less intense rainfalls
falling on saturated, easily erodible soils.
According to the Glasod assessment, in
Europe, excluding the Russian Federation,
about 114 million ha or more than 17 % of
the total land area is affected by soil erosion,
of which more than 24 million ha or
approximately 4 % show high or extreme
degradation and nearly 70 million ha or 11 %
are affected by moderate degradation.
The various regions of Europe show different
patterns, for example in the EU and EFTA
countries the area subjected to soil erosion is
about 9 % of the total land area. It increases
to 26 % in the candidate countries and to
32 % in the rest of Europe (excluding the
Russian Federation). However, these findings
are based on fragmented and nonstandardised information and hence may not
be consistent.

Soil erosion: a priority at the European


level
In April 2002, the European Commission
adopted a communication on soil protection,
endorsed by the Council of Ministers in June
2002. The communication considers soil
erosion as one of the major threats to
Europes soils and a priority for action.
Increasing the awareness amongst scientists
and policy-makers about the problem of soil
degradation through erosion in Europe is
now an urgent requirement. The
identification of areas that are vulnerable to
soil erosion can be helpful for improving our
knowledge about the extent of the areas
affected and, ultimately, for developing
measures to keep the problem under control.

Hot-spot maps of soil degradation in Europe were first published in EEA, 2000 and EEA, 2001a. The results
of a EIONET review of the hot-spot analysis and maps produced are discussed in EEA, 2002b.

Assessment and reporting on soil erosion

In a long-term perspective, the


implementation of the work on indicators
discussed in this report should certainly
contribute to improving the information
basis needed to prepare, implement and
monitor a sound European strategy on soil,
in line with the priorities set down in the
sixth environmental action programme
(EAP) and the communication on soil
protection.

Policy-relevant indicators on soil


erosion
Objective and measurable criteria with
potential to compare between areas and
monitor changes over time are needed to
describe the condition and management of
soil erosion. The driving forcespressure
stateimpact and responses (DPSIR)
assessment framework in combination with
the multi-function and multi-impact (MF-MI)
approach provides a methodology for the
integrated assessment of the soil
environment, enabling the inclusion of
causeeffect relationships into policy-relevant
indicators. The application of the DPSIR
assessment framework to soil erosion is
discussed in this report.

updated Corine land cover data are used in


combination with earth observation derived
products such as the normalised difference
vegetation index (NDVI) in order to capture
seasonal variations in land cover. Existing
policies for the protection of soils and the
degree of enforcement of such policies
should also be monitored.
Regional soil erosion assessment is needed
on a European scale in order to make
objective comparisons that may provide a
basis for further environmental analysis,
economic statements or policy development.
Some methods for carrying out regional
assessments are based on the collection of
distributed field observations, others on an
assessment of factors, and combinations of
factors, which influence erosion rates, and
others primarily on a modelling approach.
None of the reviewed methods presents stateof-the-art regional soil erosion assessments.
The Glasod and hot-spot maps can be
classified as methods based on distributed
point data, while the RIVM and Corine maps
can be classified as factor- or indicator-based
maps. Other current developments are
model-based risk analysis, such as Pesera.

Workshop findings
Following the DPSIR assessment framework,
a set of soil erosion indicators have been
proposed by the EEA and are reviewed in
Part I of this report. A major difficulty in the
development of these indicators is availability
of data. The proposed pressure indicators
link to the driving force agricultural
intensification and all have in common that
they are complex and not directly linked to
the phenomenon of soil erosion. The
identified indicators of state and impact are
difficult or expensive to measure and the
data are usually not readily available.
Indicators of response are prevention and
control measures, which are rarely in place at
present.
Land cover/use and management are the
most important factors that influence soil
erosion. Some of the indicators proposed are
related to land use. These can be regarded as
a basis for assessing pressures that may result
in soil erosion but they require further
analysis and inclusion of other factors.
Human activities that affect land use and
determine land use intensity include
agriculture, infrastructure, recreation,
mining activities or forest management. It is
therefore recommended that regularly

At the workshop the following topics were


discussed: assessment and reporting
framework; regional and spatial assessment
methods for soil erosion and data availability;
and indicators for soil erosion. Indicators
should be developed according to the
following properties and procedures:
quantitative, objectively calculated, validated
against measurements and evaluated by
experts.
The formulation of suitable remediation
measures and mitigation strategies requires a
regional assessment of soil erosion; the
extent and magnitude of areas at risk is
essential to prepare soil conservation
policies. The method should combine all
four strategies of regional erosion
assessment, i.e. measured data, expert
mapping, factor (thematic) mapping and
regional modelling. Factor- and model-based
approaches offer the advantages of
repeatability and transparency. However, the
results need to be validated against
measurements and evaluated by experts so
that the models or factor approaches can be
adapted to reflect the reality.

Executive summary

Recommendations
A set of specific recommendations for the
EEA and ETC/TE was developed with the
purpose to contribute to the EEA work
programme and to the discussion at the
European level. These recommendations are
related to the general reporting and
networking mechanism, to the DPSIR
assessment framework, to the proposed
indicators by the EEA, to the explicit
incorporation of land use into soil erosion
indicators, and to the implementation of
regional erosion assessments.
In particular, since soil erosion has impacts
on several media, such as water quality,
working links should be developed with other
ETCs and specifically with the ETC on Water.
Links with other international initiatives and
with data providers should also be
maintained.
A revised scheme for soil erosion within the
DPSIR assessment framework is proposed. It
is advised to better explore the dynamics of
the factors involved in this scheme and to
undertake a stakeholder analysis on the
proposed scheme.
The area affected by erosion is an important
indicator for the state of soil erosion, and
should be complemented with an indication
of the magnitude of erosion in particular
areas. Actual soil erosion measurements,
such as those collected for the hot-spot map,
should continue to be compiled. However,

the difficulty of making truly objective


comparisons between, and often within, areas
calls for a standardised approach to record
and particularly map the observations.
Therefore, a Europe-wide monitoring
network for soil such as proposed by the EEA
(2001b) should include monitoring of soil
erosion.
A regional assessment using modelling,
expert estimates and other methods should
be developed in order to provide a general
view and identify the hot-spot areas where a
detailed soil erosion monitoring programme
should be undertaken.
The temporal and spatial patchiness of soil
erosion favours a risk analysis approach in
order to make comparisons between regions
and to complement field measurements and
observations. Modelling efforts should be
thoroughly validated against erosion
measurements, and a clear distinction should
be made between modelled erosion risk and
present-day erosion rates. A programme to
monitor soil erosion across different agroecological regions and under different land
uses should underpin both mapping
exercises and regional soil erosion risk
assessment methods. Only then a sound
approach is ensured of estimations and
mapping features that are directly validated
and compared with measurements.
Moreover, measuring campaigns may lead to
new insights and therefore to both better
mapping and risk assessments.

10

Assessment and reporting on soil erosion

1. Introduction
1.1. Scope of the report

1.2. Background

This report has been prepared by the


Katholieke Universiteit Leuven (Catholic
University of Leuven) under contract to the
EEA and is the final result of a working group
on indicators for soil erosion. The working
group was established by the EEA in order to
progress with the work on soil in the interim
period before the new ETC on Terrestrial
Environment (ETC/TE) started in July 2001.

The EEA was established by Council


Regulation EEC (No) 1210/90 in May 1990
and started its operations in Copenhagen in
July 1994. The EEA mission is to contribute
to the improvement of the environment in
Europe and to support sustainable
development through the provision of
relevant, reliable, targeted and timely
information to policy-makers and the general
public. This should enable the Community
and Member States to take the necessary
measures to protect the environment, to
assess the results of such measures and to be
supported with the necessary technical and
scientific issues. The EEA mandate is to
provide information to Community
institutions and member countries required
to frame, identify, prepare, implement and
evaluate sound and effective policies on the
environment and to ensure that the public is
properly informed.

In 2001 the EEA carried out a peer review of


its work on soil, with particular reference to
the development of policy-relevant indicators
and the identification of probable problem
areas for soil degradation (hot spots). The
review was in particular focused on work on
indicators for soil erosion and soil sealing,
and two associated technical workshops were
held in March 2001 to facilitate this review. A
separate document was prepared for the
workshop on soil sealing and the hot-spot
review (EEA, 2002b).

The EEAs main tasks are:


Soil erosion is a natural process, occurring
over geological time, and may be caused by
water or wind. Most concerns about erosion
are related to accelerated erosion, where the
natural rate has been significantly increased
by human activities such as changes in land
cover and management. This document
focuses on accelerated erosion by water.
A workshop on assessment and reporting on
soil erosion was held in Copenhagen on 27
28 March 2001. The purpose of the workshop
was to identify a set of recommendations
concerning reporting on soil erosion (as part
of the wider theme of soil degradation) that
could then be considered for inclusion in the
work programme for the ETC/TE.
The report provides the background,
analyses the work done by the EEA on soil
erosion (Part I) and summarises the
conclusions of the workshop on indicators
for soil erosion (Part II).

(2)

1. to report on the state and trends of the


environment;
2. to establish, develop and make use of the
European Environmental Information
and Observation Network (EIONET);
3. to facilitate access to data and
information supplied to, maintained and
emanating from EEA and EIONET,
together with access to other relevant
environmental information developed by
other national and international sources.
The role of the EEA, as defined by its mission
and mandate, is therefore to provide policymakers and the public with quality
information, and to do so through a range of
products and services. The agency works as a
facilitator or bridge between member
countries (2), the Community institutions (in
particular the Commission, Parliament and
Council) and other environmental
organisations and programmes to bring
together, use, make available and thereby

To date EEA membership counts 30 countries, comprising the EU-15, three EFTA countries (Iceland,
Liechtenstein, Norway) and 11 of the 13 candidate countries (Turkey is expected to join shortly).

Introduction

improve the quality of information on the


environment relevant at the European level
for policy-making and assessment. This is
done through basic activities, including the
support to national monitoring, the
gathering and storage of existing information
and currently accessible and reliable data,
the analysis and assessment of data to
produce policy-relevant information and
indicators, the reporting of results to the
policy-makers and the dissemination of
information to the general public (Envision
model, monitor to reporting MDIAR
core activities) (Gentile, 1999a).

protection for the European Union. The


programme, proposed by the European
Commission in 2001, lays down the
Community action programme for the
period 200110 in the field of the
environment.

The European Topic Centre on Soil (ETC/


S) (3) was established by the EEA in 1996 with
the objective to provide and develop
information and data on soil aspects,
covering all EEA member countries, in order
to increase the understanding of soil as a
natural resource, document soil degradation
processes and improve the level of reliable
and comparable information about
contaminated sites, thus contributing to the
development of the EEA work programme.

Moreover, given the complex nature of the


pressures weighing on soils and the need to
build a soil policy on a sound basis of data
and assessment, a thematic strategy for soil
protection is proposed ... (European
Commission, 2001).

ETC/S operated until December 1999. A new


Topic Centre on Terrestrial Environment
(ETC/TE) started operations in July 2001.
The ETC/TE is carrying out the work
initiated by the ETCs on Soil, Land Cover
and Marine and Coastal Environment
(terrestrial part of coastal environment).
On the basis of the results of the first
EIONET workshop on soil (EEA, 2001a,b)
and a wider review of the EEA work on soil
(October 1999), in the period 2000mid2001 the implementation of the work
programme progressed through three
working groups on indicators for:
soil contamination (from local and diffuse
sources);
soil sealing; and
soil erosion.
This report is the final product of the
working group on soil erosion.

1.3. Policy developments


Since 2001 important progress took place at
the policy level. In fact, the sixth
environmental action programme (6EAP)
has introduced a new strategy on soil

(3)

The 6EAP recognises that little attention has


so far been given to soils in terms of data
collection and research. Yet, the growing
concerns on soil erosion and loss to
development as well as soil pollution
illustrate the need for a systematic approach
to soil protection ....

In April 2002, the Commission adopted a


communication on soil protection, endorsed
by the Council of Ministers in June 2002. The
communication considers soil erosion as one
of the major threats to Europes soils and a
priority for action.
A communication on soil erosion, soil
organic matter decline and soil
contamination, containing detailed
recommendations for future measures and
action, has been planned. To facilitate this
process, a conference on soil erosion and
organic matter decline in the Mediterranean
with the participation of the major
stakeholders is being organised by the
Commission and expected to take place in
2003 (European Commission, 2002).
In a long-term perspective, the
implementation of the work on indicators
discussed in this report would certainly
contribute to improving the information
basis needed to prepare, implement and
monitor a sound European strategy on soil,
in line with the priorities set down in the
6EAP and the communication on soil
protection.

1.4. Objectives and methodology of


the review
The specific objectives of this report are the
following:

ETCs are consortia of organisations that are assigned to carry out specific tasks concerning an environmental
theme. They help the EEA develop its multi-annual and annual working programmes.

11

12

Assessment and reporting on soil erosion

provide a summary overview of EEA work


on soil erosion indicators;
review the EEA European framework for
the assessment and monitoring of soil and
the proposed soil erosion indicators in
relation to data availability and analytical
soundness;
discuss the link between soil erosion
indicators and land use or land use
intensity;
review methods for assessing soil erosion on
a regional scale;
present options for future development
with particular reference to existing
European data sources; and
present the results of the workshop on
indicators on soil erosion.
The methodology adopted in the review
process consisted first of all in the evaluation
of EEA work carried out by a group of
experts and the preparation of a background
report (included in Part I). An analysis of
existing approaches for a regional assessment
of the extent of soil erosion in Europe was
also carried out (see Section 4). A selection
of national experts was asked to evaluate the
results of EEA work on soil erosion and
invited to discuss the results of the evaluation
at the workshop. Questions to guide the
review were provided (see Annex II). The
main items of the discussion and the
conclusion of the workshop are summarised
in Part II.

1.5. Soil erosion in Europe


The main problems for soils in the European
Union are irreversible losses due to
increasing soil sealing and soil erosion, and
continuing deterioration due to local and
diffuse contamination. It is envisaged that
Europes soil resource will continue to
deteriorate, probably as a result of changes in
climate, land use and other human activities.
A policy framework is needed which
recognises the environmental importance of
soil, takes account of problems arising from
the competition among its concurrent uses,
both ecological and socioeconomic, and is
aimed at maintaining its multiple functions
(EEA, 2000).
Soil erosion, in particular, is regarded as one
of the major and most widespread forms of
land degradation, and, as such, poses severe
limitations to sustainable agricultural land
use. Erosion reduces on-farm soil
productivity and contributes to water quality
problems as it causes the accumulation of

sediments and agro-chemicals in waterways.


The dynamic relationship between
agriculture and the environment requires
that erosion processes be quantified at
different scales to monitor and evaluate the
impact of agriculture and land use policies.
In Europe, soil erosion is caused mainly by
water and, to a lesser extent, by wind.
Prolonged erosion causes irreversible soil loss
over time, reducing the ecological functions
of soil: mainly biomass production, crop
yields due to removal of nutrients for plant
growth and reduction in soil filtering
capacity due to disturbance of the
hydrological cycle (from precipitation to
runoff). The major reasons are unsustainable
agricultural practices and overgrazing in
medium- and high-risk areas of land
degradation (EEA, 1999a), together with
deforestation and construction activities
(Yassoglou et al., 1998).
Soil losses are high in southern Europe, but
soil erosion due to water is becoming an
increasing problem in other parts of Europe
(EEA, 2000). Box 1 provides an overview of
the extent of soil degradation in Europe.
Some of the findings are shown in Table 1.1,
but the figures shown are only a rough
approximation of the area affected by soil
degradation.
However, Table 1.1 does indicate the
importance of water erosion in Europe in
terms of area affected. The most dominant
effect is the loss of topsoil, which is often not
conspicuous but nevertheless potentially very
damaging since it affects the most fertile part
of the soil profile. Physical factors such as
climate, topography and soil characteristics
are important in the process of soil erosion.
In part, this explains the difference between
the severe water erosion problem in Iceland
and the much less severe erosion in
Scandinavia where the climate is less harsh
and the soils are less erodible (Fournier,
1972).
The Mediterranean region is considered to
be particularly prone to erosion. This is
because it is subject to long dry periods
followed by heavy bursts of intensive rainfall,
falling on steep slopes with fragile soils and
low vegetation cover. According to presentday information (EEA, 2000, 2001), soil
erosion in north-west Europe is considered to
be slight because rain is falling on mainly
gentle slopes, is evenly distributed
throughout the year and events are less

Introduction

intensive. Consequently, the area affected by


erosion in northern Europe is much more
restricted in its extent than in southern
Europe. However, these findings are based
on fragmentised and non-standardised
information.
In parts of the Mediterranean region, erosion
has reached a stage of irreversibility and in
some places erosion has practically ceased
because there is no more soil left. In the most
extreme cases, soil erosion leads to
desertification. With a very slow rate of soil
formation, any soil loss of more than 1 t/ha/
year can be considered as irreversible within
a time span of 50100 years (EEA, 1999a).
Losses of 20 to 40 t/ha in individual storms,
that may happen once every two or three
years, are measured regularly in Europe with
losses of more than 100 t/ha in extreme
events (Morgan, 1992). It may take some
time before the effects of such erosion
become noticeable, especially in areas with
the deepest and most fertile soils or on
heavily fertilised land. However, this is all the

more dangerous because, once the effects


have become obvious, it is usually too late to
take remedial steps.
Increasing awareness amongst scientists and
policy-makers about the problem of soil
degradation through erosion in Europe is
now an urgent requirement. The
identification of areas that are vulnerable to
soil erosion can be helpful for improving our
knowledge about the extent of the areas
affected and, ultimately, for developing
measures to keep the problem under control.
Attention is focused mainly on rill- and
interrill erosion because this type of erosion
affects the largest area. Other forms of
erosion are also important, for example,
gully erosion, landslides and, to a lesser
extent, wind erosion. Some of these,
particularly gully erosion and landslides, have
serious consequences for land use systems
and populations, but in overall terms are still
relatively localised (see Annex IV for a
description of the different types of erosion).

Box 1 Soil erosion in Europe


Table 1.1

Extent of human-induced soil degradation by erosion in Europe (million hectares)


Erosion type
Accession countries

EFTA countries

Rest of Europe

High

Extreme

Total

4.5

29.2

14.7

0.0

48.4

Wind erosion

0.0

0.0

0.0

0.0

0.0

AC total

4.5

29.2

14.7

0.0

48.4

Water erosion

0.8

1.5

0.0

0.0

2.3

Wind erosion

0.6

1.3

0.0

0.0

1.9

EF total

1.3

2.9

0.0

0.0

4.2

Water erosion

0.8

19.3

6.5

1.0

27.7

Wind erosion

0.0

5.8

0.0

0.7

6.5

0.8

25.1

6.5

1.7

34.2

12.8

11.9

1.4

0.0

26.2

Water erosion
Wind erosion

Europe (excl. the


Russian Federation)

Moderate

Water erosion

ER total
European Union

Light

1.0

0.1

0.0

0.0

1.1

EU total

13.8

12.0

1.4

0.0

27.3

Water erosion

18.9

62.0

22.6

1.1

104.6

Wind erosion

1.6

7.2

0.0

0.7

9.5

20.5

69.2

22.6

1.8

114.1

All Europe total

(17.4 % of
total land area)

Note: Any mismatch between totals and disaggregated figures is due to the rounding process.
Source: EEA data elaboration from Glasod (Oldeman, 1991; Van Lynden, 1995; data: UNEP and ISRIC
through UNEP/GRID Geneva, 2001).
According to the Glasod assessment, in Europe, excluding the Russian Federation, about 114 million ha or
more than 17 % of the total land area is affected by soil erosion, of which more than 24 million ha or
approximately 4 % show high or extreme degradation and nearly 70 million ha or 11 % are affected by
moderate degradation. The major type of degradation is erosion by water (about 16 % of the total land area),
while erosion by wind interests only 1.5 % of the territory.
The various regions of Europe show different patterns, for example in the EU and EFTA countries the area
subjected to soil erosion is about 9 % of the total land area. It increases to 26 % in the candidates countries
and to 32 % in the rest of Europe (excluding the Russian Federation).

13

14

Assessment and reporting on soil erosion

Part I Assessment and reporting


on soil erosion
2. A European framework for the
assessment and monitoring of soil
The degradation of the environment
through soil erosion is an important concern
for policy-makers.
Objective and measurable criteria with the
potential to compare between areas and
monitor changes over time are needed to
describe the condition and management of
land resources and the pressures exerted
upon the land.
There is now a requirement for
environmental protection agencies to
periodically report on the state of the
environment and particularly whether this is
deteriorating, stable or improving. Agencies
are dealing more commonly with a
degrading environment, hence the search for
indicators that can quantify this
degradation in some way.
International organisations such as the EEA,
OECD and UN have initiated programmes
on developing measurable and policyrelevant agri-environmental indicators to
assess and monitor progress in reaching
sustainable development, as defined in
Agenda 21 by the United Nations
Conference on Environment and
Development (UNCED).

2.1. The assessment framework


An update of the state of progress of the EEA
soil work programme and the relevance of
indicator development including the
reporting system were presented at the
EIONET workshop on indicators for soil
contamination in Vienna, 1819 January
2001 (EEA, 2002a, b).
The concept of multiple soil functions and
competition is crucial in understanding
current soil protection problems and their
multiple impacts on the environment. The
EEA considers soil with its multiple
ecological and socioeconomic functions and
multiple impacts as having a fundamental
role in Europes environment (EEA, 1999a).
The ecological functions comprise

production of biomass; filtering, buffering


and transforming; gene reserve and
protection of flora and fauna. The
socioeconomic functions include support to
human settlements; source of raw materials,
including water; and protection and
preservation of cultural heritage. Soil
degradation means loss or deterioration of its
functions (Blum, 1998). Soil losses due to
erosion can be considered as irreversible in
relation to the time needed for soil to form
or regenerate itself.
The OECD DSR framework (driving force
stateresponse) has established a holistic
systems approach to include cause-effect
relationships (OECD, 1993). The OECD
model has been extended by the EEA to
cover the causes (pressures) and the impacts
on the environment (EEA, 1999b, 2000).
The DPSIR assessment framework shows a
chain of causeseffects from driving forces
(activities) to pressures, to changes on the
state of environment, to impacts and
responses (EEA, 1999, 2000). DPSIR is based
on the assumption that economic activities
and societys behaviour affect environmental
quality. The relationships between these
phenomena can be complex. DPSIR
highlights the connection between the causes
of environmental problems, their impacts
and societys response to them, in an
integrated way. The DPSIR applied to soil
resources is shown in Figure 2.1.
In addition to the DPSIR, the EEA has
defined the multi-function and multi-impact
approach (MF/MI), based on the
recognition of the role played by the soil
multiple functions and the problems arising
from the competition between these
functions (see Figure 2.2).
Both DPSIR and MF/MI are analytical tools
for the definition of policy-relevant indicators
to describe pressures placed upon soil
resources, changes in the state of soil, and
impacts or responses by society to these
changes, within the context of policy and soil

Part I Assessment and reporting on soil erosion

The DPSIR assessment framework applied to soil (EEA 2000)

SECONDARY PROTECTION

PRIMARY PROTECTION
Desertification Convention
Development of a European
soil protection policy
Human population
Land development
Tourism
Agriculture
Transport
Industry/Energy
Mining
Natural events
Climate change
Water stress

Responses

Figure 2.1
Source: EEA, 1999a.

CAP reform
Nitrate directive
Sewage sludge directive
Water framework directive
Air pollution prevention measures
Spatial development/Land use
measures (EIA;ESDP)

Driving
Forces

Emissions to air, water


and land.
Land consumption
Agricultural intensification and
management practices
Forest fires

15

INDIRECT(effects on other

Pressures

Impact

media, ecosystems and


human population)
Changes in population
size and distribution
Human health
Change of biodiversity (soil
habitats and species)
Plant toxicity
Changes in crop yields
Changes in forest health
and productivity
Contamination of surface
and groundwater
Climate change
Water stress

DIRECT(Changes in soil

State

function)

SOIL DEGRADATION
SOIL LOSS

Local and diffuse contamination


Soil acidification
Salinisation
Nutrient load (soil eutrophication)
Physical deterioration

Soil sealing
Soil erosion
Large scale land movement

Examples from the multi-function/multi-impact approach

Figure 2.2
Source: EEA, 1999a.

Acidification
Change of
biodiversity
Preservation of
cultural heritage

Water stress

Filtering/
Buffering

Biomass
production

Soil

Species gene
reserve and
protection

Source of
raw material

Climate
change

Support to human
settlements

Examples of multi-impact approach


pressure on soil / impact on soil functions
impact of loss / deterioration of soil functions

resource management (Gentile, 1999a).


These tools also provide a framework for the
subsequent interpretation and assessment of
the indicators.

of a phenomenon/environment/area with
significance extending beyond that directly
associated with a parameter value (OECD,
1993).

In environmental monitoring, indicators


have been defined as parameters, or values
derived from parameters, which point to/
provide information about/describe the state

OECD (1993, 1999) defines agrienvironmental indicators (AEIs) as attributes


of land units, which are:

16

Assessment and reporting on soil erosion

Figure 2.3

The EEA information strategy 'from national monitoring to European reporting' (MDIAR framework)

Source: EEA, 2001b.

policy relevant and have utility for users; i.e.


the AEIs should:
provide a representative picture of
environmental conditions, pressures on
the environment or societys responses;
be simple, easy to interpret and able to
show trends over time;
be responsive to changes in the
environment and related human
activities;
provide a basis for international
comparisons;
be either national in scope or applicable
to regional environmental issues of
national significance;
have a threshold or reference value
against which to compare them so that
users are able to assess the significance of
the values associated with them;
analytically sound; i.e. the AEIs should:
be theoretically well founded in technical
and scientific terms;
be based on international standards and
international consensus about their
validity;
lend themselves to being linked to
economic models, forecasting and
information systems;
measurable; i.e. the data required to
support the AEIs should be:
readily available or made available at a
reasonable cost/benefit ratio;
adequately documented and of known
quality;

updated at regular intervals in


accordance with reliable procedures.
In addition to the above criteria, the EEA
selects indicators having in mind the target
audience, together with the most suitable
level of aggregation and the availability of
data needed to compile them (Gentile,
1999a). An overview of the situation is
provided by indicators with a high level of
aggregation, so-called headline indicators
(Gentile, 1999a), while detailed indicators
are needed to better understand underlying
trends or existing links between policy
measures and their effects. The challenge is
finding an appropriate balance between
simplification and completeness.
The EEA together with its EIONET partners,
including the European Topic Centres
(ETCs), are facilitating the process from
national monitoring to European reporting
(Figure 2.3). The MDIAR framework consists
of monitoring, data collection, information,
assessment and reporting. The set up of a
European soil-monitoring network
harmonises national networks and enables
data comparability. Data flow and
management entails organisation and
storage in databases. Data are integrated into
indicators and assessed using the DPSIR and
MF/MI approaches. Reporting enables
communication of the results obtained. The
MDIAR chain concentrates on matching the
best available environmental information
with the best needed environmental and
economic information.

Part I Assessment and reporting on soil erosion

The DPSIR assessment framework applied to soil erosion

Good agricultural practices

Desertification Convention

- Land use practices in accordance


with sustainable development
- Local programmes on soil
erosion consulting

Development of a European
soil protection policy
(Human population)
Land development
Natural events
Agriculture*
*Intensification

(De-forestation)
(Forest fires)
Land use practices

Responses
Driving
Forces
On-site
Loss of soil fertility
Changes in soil functions
Changes in crop yields
Desertification

Pressures

Impact
Off-site
Effects on other media, e.g.
- water stress
- eutrophication

State
On-site: SOIL DEGRADATION
Physical deterioration
SOIL LOSS
Off-site: emission to air, water and land

2.2. The DPSIR assessment


framework applied to soil
erosion
Figure 2.4 presents the DPSIR assessment
framework applied to soil erosion as
proposed by EEA-ETC/S (1999). Possible
driving forces can be grouped according to
human activity and physical phenomena,
which in turn result in potential pressures on
the land. An important driving force related
to soil erosion is the intensification of
agriculture. Intensification of agriculture
encourages unsustainable land use practices
and deforestation, which in turn enhance the
risk of soil erosion. These pressures may
change the state of the soil resources, and
result in soil loss. Soil loss is recognised to
have both direct and indirect impacts on the
environment, expressed in terms of on-site
and off-site effects, respectively (Figure 3.4).
The responses at the European level include
CAP reform, soil conservation measures and
land use practices in accordance with
sustainable development. However, a
European policy framework on soil
protection, similar to those already in place
for air and water, does not exist. Moreover,
there is no reporting mechanism in place to
assess whether existing measures are leading
to improvement of soil conditions or to
(4)

Economic aspects, e.g.


- impediment of traffic
- disturbance of drainage
Changes in soil functions
Changes in crop yields

gauge the level of implementation of existing


legislation (EEA, 2000) (4).
The assessment carried out through the
DPSIR assessment framework does not aim at
understanding or analysing soil erosion as a
process, but provides information to support
policy-makers actions so that the necessary
measures can be defined and the effect of
current measures can be assessed.

2.3. Is the proposed DPSIR


assessment framework adequate
to comprehend soil erosion?
The result of the application of the DPSIR
and MF/MI assessment tools to soil erosion is
the identification of a set of policy-relevant
indicators. However, it has to be recognised
that there is a huge difference between actual
and potential soil erosion, which is not
adequately reflected in the present
framework (EEA-ETC/S, 1999). Indicators
describing the driving forces and pressures
may affect the risk of soil erosion, but they
may not affect soil erosion in itself, which
also depends on physical parameters such as
climate and relief. A mechanism is therefore
needed to jointly estimate the potential and
actual risk, based on links between the
identified driving force and pressure

In April 2002, the Commission adopted a communication on soil protection, later endorsed by the Council of
Ministers in June 2002. The communication considers soil erosion as one of the major threats to Europes soil
and a priority for action (European Commission 2002; see also Section 1.4).

17

Figure 2.4
Source: EEA-ETC/S,
1999.

18

Assessment and reporting on soil erosion

Figure 2.5

The DPSIR assessment framework applied to soil erosion modified from EEA, 2000, and EEA-ETC/S, 1999
PRIMARY PROTECTION
Development of a European
soil protection policy
Human population
Land development
Tourism
Agriculture
Transport
Natural events
Climate change

Land cover changes


Precipitation

Responses

SECONDARY PROTECTION
CAP-reform
Spatial development/Land use
measures (EIA; ESDP)

Driving
Forces

Pressures

Impact

State
SOIL LOSS
Soil erosion
Mass movement
Change in soil quality (depth)

indicators, and on an estimation or


measurement of what is actually happening.
Agricultural intensification is seen as the
most important driving force (EEA-ETC/S,
1999; EEA, 2000). However, tourism and
transport could be added to the list of driving
forces. The effect they have in common is
that they change the land cover, which is the
major pressure indicator for soil erosion.
This would lead to a revised scheme for soil
erosion within the DPSIR assessment
framework, presented in Figure 2.5.
The DPSIR assessment framework lends itself
to systems analysis and as such is very useful
in describing the relationships between the
origins and consequences of environmental
problems. Obviously, the real world is more
complex than can be expressed in simple
causal relationships. Linkages between the
different types of indicators are explored
through the DPSIR chain. However, the
linkages deserve further attention, not least
to capture the dynamics of the system.
Moreover, linkages within one type of
indicators (e.g. pressures) are not explored,
despite their repeatedly reported
importance.
The emphasis of the DPSIR assessment
framework is on socioeconomic related
indicators, while physical indicators of
pressure are not fully explored, nor explicitly
mentioned. Climate change is considered as

DIRECT
(Changes in soil function)
Loss of soil fertility
Contamination of surface water

INDIRECT
(effects on other media,
ecosystems and human population)
Changes in population
size and distribution
Change of biodiversity (soil
habitats and species)
Changes in crop yields
Desertification
Water stress

a driving force but only in the sense that it


relates to human activities. Important
physical factors that influence soil erosion
are topography, soil type, soil vulnerability
and climatic factors (particularly rainfall).
These factors cannot be separated from the
identified pressure indicators. On the other
hand, they are implicitly incorporated into
indicators of state.
A major problem with soil erosion is the
temporal and spatial scale of reporting and
the spatial extent to which the phenomenon
occurs. Although problems of both spatial
and temporal patchiness are well recognised
in the various reports (EEA, 2000; EEA,
2001a), a more integrated approach of
reporting seems recommendable. One
solution could be to develop a regional
model that allows for estimating the potential
soil erosion risk, combined with periodical
monitoring of actual soil erosion in selected
test areas. The regional soil erosion model
should express the links between the
different biophysical and socioeconomic
factors, i.e. be process-based; establish various
spatial and temporal resolution linkages; and
provide a nested strategy of focusing on
environmentally sensitive areas which may
require remedial measures to be taken.
Sections 3 and 5 provide more details on the
requirements for future regional soil erosion
reporting in order to develop sound
indicators of state.

Part I Assessment and reporting on soil erosion

In the different reports made by the EEA, it is


recognised that a distinction ought to be
made between on-site and off-site impacts of
soil erosion. This distinction, however,
already applies at an earlier stage in the
DPSIR chain, namely at the stage of state
indicators. Soil erosion can be measured in
terms of actual sediment loss per unit area
(on site) or in terms of sediment delivery into
streams or rivers (off site).
The current level of detail chosen for the
application of the DPSIR assessment
framework to soil erosion implicitly enables
the identification of broad groups of actors
related to the perceived environmental
problem. However, the full identification of
the several actors involved requires a more
detailed stakeholder analysis. Environmental
problems can be identified and discussed by
each group of stakeholders using
participatory methods for eliciting the
various aspects of the perceived problem. A
general stakeholder analysis ultimately helps
formulating policies for remediation and
mitigation strategies.
In conclusion, the DPSIR assessment
framework is an excellent tool onto which
further extensions and strategies of reporting
can be built. The framework sets a good basis
for identifying the different factors
influencing soil erosion, and should be
coupled with a detailed stakeholder analysis
in order to identify the full range of actors in
the DPSIR chain.

2.4. EEA typology of indicators


applied to soil erosion
The EEA identifies four different types of
indicators (EEA, 1999b):
descriptive indicators, describing the actual
situation in the DPSIR assessment
framework;
performance indicators, comparing the
actual situation with a specific set of
desirable conditions in terms of a distance
to target assessment;
efficiency indicators, expressing the
relation between separate elements of the
causal chain such as between
environmental pressures and human
activities;
total welfare indicators, measuring
sustainability in the form of an index
(Green GDP or index of sustainable

economic welfare), currently not within the


EEAs mandate.
Efforts related to soil erosion have
concentrated on descriptive indicators within
the DPSIR philosophy. Without a European
policy framework on soil protection, however,
little progress can be expected on the other
three types of indicators. Sound advice on
how to develop performance indicators on
soil protection will be one of the challenges
of the European Topic Centre on Terrestrial
Environment.

Indicators of soil erosion


and data availability
2.5. Introduction
The development of policy-relevant
indicators for soil was one of the main
activities of the European Topic Centre on
Soil (ETC/S). The EEA has proposed and
discussed a set of indicators for soil erosion
(EEA-ETC/S, 1999). ETC/S work aimed to
identify policy-relevant indicators for soil
erosion and to update the existing databases
by means of data collection requests. Further
recommendations were made to assess data
needs and availability, and to set up
monitoring activities. Since climate, soil and
relief are fairly static variables, the ETC/S
recommended ground cover measurements
to be closely monitored. The EEA have
drawn up a list of policy-relevant indicators
for soil (Gentile, 1999b), which was
presented at the EIONET workshop in
October 1999 and at the first Soil Forum held
in Berlin in November 1999 (Table 2.1)
(EEA, 2001a, b).

2.6. Review
Indicators for soil erosion should incorporate
the following characteristics.
The indicators will be a measure of soil loss
due to erosion as a result of climate,
topography, soil properties, land cover and
land management.
The extent and severity of both potential
and actual soil erosion risk will have to be
quantified and related to land cover
changes.
The nature of soil erosion has to be assessed
in order to evaluate the on-site loss and the
possible off-site impacts.

19

20

Assessment and reporting on soil erosion

Table 2.1
Source: EEA-ETC/S,
1999; Gentile, 1999b.

EEA draft list of policy-relevant indicators for soil


Issue / question

Indicator

Units

Intensity of agriculture:
Degree of agricultural land use (ALU)?
To what extent does
ALU intensify during
a specified time within a given country?

DPS
IR

Soil
degrad
ation
pattern

Shortterm
core
indicat
ors

Comment

Not
applicab
le

Yes

Index of
output vs.
input

Consumption of fertilisers
per defined region (e.g.
Member State) (and its increase)

t/ha

Soil
erosion

No

Available
in Eurostat
and OECD

Average farm size per defined region (e.g. Member


State) (and its increase)

Euro/ha

D/P

Soil
erosion

No

Low
priority

Average field sizes (and its


increase)

Euro/ha

D/P

Soil
erosion

No

Low
priority

Average crop yield per area


(and its increase)

t/ha

D/P

Soil
erosion

No

Desirable
but not
key

Average net profit per area

Euro/ha
yr

Soil
erosion

No

Low
priority

Number of grazing animals

No/ha

Soil
erosion

No

Desirable
but not
key

To what extent is the


area of member
countries affected by
soil erosion (both
wind and water erosion)?

Short term: rough estimations by the countries:


percentage of area affected
by soil erosion per defined
region (e.g. Member State)

Soil
erosion

No

Desirable
but
difficult to
obtain

To what extent is the


total area of Europe
affected by soil erosion (both wind and
water erosion)?

Depending on the progress


of validation of the ISRIC
map

km2

Soil
erosion

No

Also
outlooks

What is the extent of


total soil loss by soil
erosion (water erosion)?

Short term: rough estimations:


estimation of the total gross
erosion of defined areas
based on the sediment delivery ratio of selected rivers
(in dependence of the watershed area)

Soil
erosion

Yes

Also
outlooks

What is being done


to remove off-site
damages by soil erosion?

Expenditures for removals


of sediment deposits in
built-up areas (traffic routes,
houses)

Euro

I/R

Soil
erosion

Yes

Desirable
but not
key

How much is spent


on sustainable farming?

Local agricultural programmes to enforce sustainable farming management


systems (incl. terminated
set-aside of arable land)

Euro

Soil
erosion/
Diffuse
contami
nation

No

Desirable
but not
key

How much is spent


on erosion prevention?

Expenditures for special soil


erosion prevention programs, forest fire protection

Euro

Soil
erosion

No

Desirable
but not
key

To what extent is the


erosion risk area of
member countries
protected from soil
erosion (both wind
and water erosion)?

Portion of actual erosion risk


area under erosion control
management (set-aside arable land, strip cropping,
contour ploughing, crop
changing, balanced grazing,
reforested), on total area of
actual erosion risk

Soil
erosion

No

Key but
difficult

Note: Priority indicators are marked in bold.

Yes

Yes

Yes

Yes

Yes

Yes

Yes

No

Yes

Yes

Yes

Farm size and trend

Field size and trend

Crop yield and trend

Net profit and trend

Stocking rate and


trend

Actual soil erosion

Delivery of sediment

Removal of sediment

Prevention
(agriculture)

Prevention (forest,
natural)

Erosion control

Yes

No

No

No

No

No

No

No

No

In part

Yes

No

Easy to
interpret

Representative

Yes

Utility

Policy
relevant

Fertiliser use and


trend

EEA Indicator

Yes

Yes

Yes

No

???

???

Yes

Yes

Yes

Yes

Sometime
s

Probably

Comparable

Yes

Yes

Yes

No

Yes

Yes

No

Probably

Yes

Probably

Probably

???

Scientific/
Technically

Rarely
available

No

Probably
not

???

Difficult to
measure

Rarely
available

National/
EU

National

National/
EU

National/
regional

Nationally

Eurostat,
OECD

Data
available

Analytical soundness

In part

In part

In part

No

In part

In part

Yes

Yes

Yes

Yes

Yes

Yes

Documented

Measurability

No

No

No

No

No

No

Yes

Yes

Yes

Yes

Periodically

Yes

Updated

Direct

Direct

Direct

Direct

Direct

Direct

Complex

Not relevant

Complex

Complex

Complex

Complex

Effect

Usually piecemeal

Usually piecemeal

Usually piecemeal

Comprehensive measurements not possible

Measurement difficult, source difficult to establish

Extent not known, expensive to measure

Dichotomy between intensive indoor and outdoor


stocking

Data for actual and estimated (CGMS) yields

Data partially available

Not linked directly

Economic criterion, link variable

Comments

Part I Assessment and reporting on soil erosion

EEA indicators for soil erosion tested according to the OECD criteria

21

Table 2.2

22

Assessment and reporting on soil erosion

As accelerated erosion is a complex process,


it is necessary to develop indicators that
identify the causes. Physical factors that
influence erosion rates include topography,
soils, climate and land cover. Land cover is in
turn influenced by the socioeconomic
environment and as such by anthropogenic
activities, notably land use and management.

Section 4 concentrates on other aspects


related to pressure indicators. One major
remark is that the intensity of agriculture
should never be evaluated alone in relation
to erosion. Soil loss due to erosion is a result
of climate, topography, soil properties, land
cover and land management. Land cover also
includes the natural vegetation.

Table 2.2 lists the EEA indicators for soil


erosion with brief comments on the OECD
criteria listed in Section 2.1. The first six
indicators relate to pressures as a result of
agricultural intensification. These pressure
indicators all have in common that they are
complex and not directly linked to the
phenomenon of soil erosion. The identified
indicators of state and impact are difficult or
expensive to measure and the data are
usually not readily available. Indicators of
response are prevention and control
measures, which are rarely in existence at
present. A more comprehensive discussion
follows in the next sections.

2.6.1.1. Consumption of fertilisers


The proposed indicator is the consumption
of fertilisers per defined region (e.g. Member
State), measured in tonnes/ha. The
consumption of fertilisers can give an
indication of the intensification of
agriculture (EEA-ETC/S, 1999). Another
positive aspect is that data on estimated
consumption of fertilisers are available at
national level from the European Fertiliser
Manufacturers Association (EFMA) or via
Eurostat/OECD.

2.6.1. Indicators of driving forces and


pressures
According to the EEA (EEA-ETC/S, 1999),
the main driving force on soil that causes
erosion in regions with potential and actual
soil erosion risks is the intensification of
agriculture. This is a complex indicator and it
is related to different pressure indicators.
The corresponding pressures are costeffective but unsustainable land use practices,
the use of machinery for the cultivation of
enlarged fields, the overgrazing and other
instruments of intensive land use practices
(EEA-ETC/S, 1999). Average field sizes (and
increase of field sizes), combined with
average farm size per region as well as the
consumption of fertilisers and the number of
grazing animals, give an indication of the
intensification of agriculture.
The intensification of agriculture is not
necessarily directly related to soil erosion.
The higher the degree of intensity of
agricultural land use the higher may be the
soil loss by water and wind erosion in
potentially high erosion risk areas, but the
reverse could equally be true. For example an
intensive farming system employing soil
conservation measures, such as terracing and
cover crops, may result in less soil erosion
than a more extensive system that does not
involve conservation techniques. Intensive
land use can be combined with efficient soil
conservation measures.

The reliability of the data used to calculate


this indicator may be seriously questioned.
The main source of information on fertilisers
in Europe is EFMA (see http://
www.efma.org/). The data from EFMA are
the production of fertiliser from the
associated members. Then the EFMA uses
data on imports and exports to calculate
fertiliser use or consumption at the national
level. For example, the current approach is:
(Fertiliser consumption {in a Member State})
= (production) (exports) + (imports)
(2.1)
To determine the actual fertiliser use by
equation (2.1), certain adjustments should
be applied to take account of losses (e.g. 10
15 %) and use outside general agriculture,
for example in market and domestic gardens
(e.g. 10 %). However, fertiliser applications
vary for different crops so it is not possible to
predict the consumption of fertilisers using
this approach without knowing precisely the
spatial distribution of crops and local
agricultural practices.
The main conclusion is the higher the
degree of intensity of agricultural land use,
the higher the likely loss of soil through
water and wind erosion in potentially high
erosion risk areas (EEA-ETC/S, 1999).
However, fertiliser consumption data cannot
be determined accurately enough to be used
as an indicator for soil erosion at the scale
required. Moreover, fertiliser applications
may increase when using soil conservation
measures so that soil erosion decreases.

Part I Assessment and reporting on soil erosion

Together with consumption of fertilisers,


average farm size (per defined region (e.g.
Member State) and its increase), average
field size (and its increase), average crop
yield (per area and its increase), average net
profit (per area) and number of grazing
animals give an indication of the
intensification of agriculture (EEA-ETC/S,
1999). This does not necessarily imply an
increase in soil erosion. The factors that
relate directly to erosion are soil type,
topography, crop cover and precipitation.
However, knowing the contribution of the
agro-economic sectors to soil erosion is
essential for the policy-makers to be able to
take the requisite measures and monitor
their implementation, but the lack of good
quality data hinders the development of
suitable indicators in the short term.
2.6.1.2. Average farm and field size
The proposed indicators are average farm
size per defined region (e.g. Member State)
(and its increase) and average field size
(and its increase), both measured in ha.
Data on farm and field size are available at
national and European level. These data are
periodically updated, with full farm surveys
every 10 years and sample surveys of farm
structure every two to four years. However,
these data are only averages on a large area
basis, e.g. Member State, and there is no
demonstrable direct link between actual soil
erosion and either farm or field size.
As stated above, average farm size and
average field size can give an indication of
the intensification of agriculture.
Furthermore, monitoring an increase in farm
and field size would infer increased
intensification of agricultural practices.
However, only in areas of medium and high
erosion risk would increased soil loss be likely
to result directly from an intensification of
agriculture. As indicators of soil erosion,
these parameters cannot be used
independently.
2.6.1.3. Crop yields and animal numbers
The proposed indicators are average crop
yield per area and its increase, measured in
tonnes per ha and number of grazing
animals in numbers per ha. Data on crop
yields and stock numbers are updated
annually at national and European level.
Where there are difficulties in obtaining crop
yield data, forecasts are available from
simulation modelling, for example the crop
growth monitoring system (CGMS) that
underpins the MARS project.

However, these data are usually only averages


on a large area basis (e.g. national level).
Moreover, there is no demonstrable direct
link between actual soil erosion and either
crop yields or animal numbers. Crop types
and the extent of the area devoted to each
crop type also influence soil erosion.
Interpreting the data on grazing animals is
confounded by the type of land use system
(for example, livestock fed by fodder or
outdoor grazing).
As stated above for average farm size and
average field size, crop yields and animal
numbers can only give an indication of the
state of intensification of agriculture.
Monitoring increases in yields and/or
stocking densities would infer increased
intensification of agricultural practices.
However, the potentially adverse effects
might only occur in areas of medium and
high erosion risk and could be mitigated by
the adoption of conservation techniques.
Therefore, as indicators of soil erosion, these
parameters cannot be used independently.
2.6.2. Indicators of state
In the absence of direct measurements, soil
erosion state indicators should be able to
provide a picture of both the extent and the
severity of the potential/actual soil erosion
risk (EEA-ETC/S, 1999). The potential risk
calculations should take into account
climatic, topographic and edaphic
conditions, whereas the actual risk should
take into account both vegetation cover and
actual land use. The comparison of the
potential with actual soil erosion risk could
be considered as a risk due to land use
changes and practices.
The indicators of state should also provide
information on the rate of the actual soil loss
under the existing soil management and
erosion control practices and on the rate of
soil loss tolerance.
2.6.2.1. Area affected by soil erosion
Two indicators are proposed as measures of
the area affected by erosion: percentage of
area affected by soil erosion per defined
region (e.g. rough estimations by the
Member States) and extent (km) to which
the total area is affected by soil erosion (both
wind and water erosion).
The area affected by erosion is an important
indicator for soil erosion. Trends in soil
erosion could be established from periodic
estimates. A number of national databases

23

24

Assessment and reporting on soil erosion

are available for making estimates at national


level. However, national databases are not
available for all EU countries. Estimates of
the area actually affected by soil erosion at
regional and national levels are not readily
available. This is because measurements of
actual erosion are difficult and usually
expensive to make or erosion is not a
problem at all. Soil erosion often takes place
surreptitiously and over long periods before
the true extent is appreciated. Accurate data
are therefore scarce.
An alternative to the direct measurements
and actual erosion estimates would be the use
of models to estimate the risk of erosion,
potential and actual (see Section 3.3).
Some estimates of actual erosion risk have
already been made: e.g. Corine and recent
work by the European Soil Bureau (Van der
Knijff et al., 1999, 2000; Kirkby and King,
1998). The European Soil Database, the
MARS meteorological database, digital
elevation data (DTM/DEM), Corine land
cover are available for making estimates at
European level.
This is one of the key indicators for soil
erosion that should be adopted by the EEA.
Estimates from Member States, based on
national data sets, could be compared with
estimates derived from European data sets
(e.g. the European Soil Database). Although,
there are difficulties in making
measurements, existing data should be
compiled and stored centrally for
comparison with model estimates. Erosion
models offer a mechanism whereby the area
affected by erosion can be estimated. An
appropriate model should be identified and
used in conjunction with standard data sets
to provide standardised estimates of the areas
at risk from soil. The result would be to
provide an appropriate state indicator
including time series for use by policymakers (5).
2.6.2.2. Actual versus potential soil loss
The proposed indicator is the extent of total
soil loss by soil erosion due to water,
measured in tonnes per ha per annum. EEAETC/S (1999) propose the USLE equation
or preferably other recent regional
quantitative models to estimate on-site soil
erosion. The comparison between the

(5)

potential and actual soil erosion risk can be


considered as a risk indicator for land use
changes.
On a medium timescale soil erosion maps
could be prepared on the basis of the Soil
geographical database of Europe (soil data),
the Soil regions of Europe map
(topographic data), land cover data (Corine
or better, more recent remote sensing
images) and climatic data. The Pesera
methodology is based on the use of these
data and will be able to result in a panEuropean soil erosion risk map (see Annex
III Workshop paper by Gobin and Kirkby).
A regional model that allows for estimating
the potential soil erosion risk should be
combined with periodical monitoring of
actual soil erosion in selected test areas.
2.6.2.3. Transport of sediments
In order to quantify actual soil losses, the
gross erosion in defined watersheds of
selected rivers could be estimated from the
sediment delivery ratio, in t per m per year.
Data on sediment transfer are proposed as a
proxy indicator of actual soil loss (EEA-ETC/
S, 1999). However, the sediment source
remains highly uncertain and can rarely be
traced back to surrounding land, riverbanks
or channel. A digital database to define
catchment boundaries in Europe (scale 1: 1
000 000) is under development at the Space
Applications Institute, JRC, in coordination
with the EEA. Data on sediment
concentrations and annual suspended
sediment yields should always be related to
the catchment area. However, the data may
not be readily available at present. The EEA
has established a European freshwater
monitoring network (EuroWaterNet), which
could be a possible source for data on river
sediments.
A difficulty to consider is that data on
sediment transport for selected rivers do not
relate to the exact source of the sediment.
The sediment loads in rivers can only give an
indication of the erosion taking place over
large areas. As an indicator for soil erosion,
sediment delivery data are rarely accurate
enough to be an independent indicator. EEAETC/S (1999), in fact, consider the transport
of sediments as an indicator of impact.

The Pesera project, funded by the European Commission under the 5thFramework Programme for Research,
aims to provide and finalise such a model within the next year and could possibly provide better estimates
(Gobin and Govers, 2001). See Annex III workshop paper by Gobin and Kirkby.

Part I Assessment and reporting on soil erosion

2.6.3. Indicators of impact


Indicators of impact could be divided into
on-site and off-site impacts. On-site impacts
in terms of loss of soil fertility are mostly
compensated for by technical advances. On
the other hand, off-site impacts are more
easily measured and could be expressed in
economic terms.
2.6.3.1. Removal of sediment deposits
The proposed indicator relates to
expenditures for removals of sediment
deposits in built-up areas (traffic routes,
houses). Data on remedial measures are
rarely available at the national level, let alone
at the European level. However, there are
subsidies provided by the EU for remedial
works via the CAP. Remedial measures usually
follow major floods and should be linked to
flood forecasting systems.
2.6.4. Indicators of response
The comparison of soil erosion rates with, yet
to be defined, soil loss tolerances for
different regions would provide estimates of
the impacts and the required response.
2.6.4.1. Conservation practices
An important indicator of response is the
expenditure for local agricultural
programmes to enforce sustainable farming
management systems (including the set-aside
of arable land). These practices include
contouring, terracing, strip cultivation, and
subsurface drainage (Renard et al., 1997).
Other measures involve adoption of
minimum tillage systems, planting cover
crops (to reduce the duration of bare
ground), and changing fundamentally the
land use system (for example, conversion
from arable to pasture).
Conservation practices have been
demonstrated to considerably reduce soil loss
through erosion in other parts of the world.
Many of these practices increase plant cover
and therefore directly reduce erosion. Many
are also recognised as good agricultural
practice. However, data and information on
conservation practices are rarely collected
systematically and stored centrally in Europe.
Conservation practices are important in
reducing or eliminating soil erosion but they
are usually only adopted after soil erosion has
been identified as a significant problem.
2.6.4.2. Mitigation strategies
The indicator proposed is the expenditures
for special soil erosion prevention

programmes, including forest fire


protection.
Measures involve implementation of fire
prevention systems and building of holding
reservoirs. Conservation practices are
important in reducing or eliminating soil
erosion but they are usually only adopted
after soil erosion has been identified as a
significant problem. Data and information
on conservation practices are rarely collected
systematically and stored centrally in Europe.

2.7. Options for the future:


determining the risk of soil
erosion
From the review of the current indicators for
soil selected by the EEA, it is concluded that,
from a scientific and technical standpoint,
the most appropriate state indicator is the
area affected by erosion. However, because
there is a serious lack of direct measurements
of soil loss, by water and by wind, a surrogate
parameter or indicator is needed.
Conventional wisdom suggests that the area
actually affected by erosion should be directly
related to the area at risk from erosion,
provided that the area at risk has been
determined using an appropriate model of
soil erosion, together with the necessary
spatial data sets. Soil erosion takes place at
the field scale, and the main problem is that
the digital data sets used to quantify the
factors causing erosion are usually too coarse
(in terms of spatial resolution) to enable
accurate estimation of soil losses at this scale.
An important surrogate indicator of actual
erosion is its risk. A risk is the chance that
some undesirable event may occur. Risk
assessment involves the identification of the
risk, and the measurement of the exposure to
that risk. The response to risk assessment may
be to initiate categorisation of the risk and/
or to introduce measures to manage the risk.
In some cases, the risk may simply be
accepted. In other cases, the priority will be
to adopt a mitigation strategy. Such risk
management, traditionally a significant
activity in the commercial sector (e.g. the
insurance industry) has now been adopted in
the environmental protection field.
Various approaches can be adopted for
assessing soil erosion risk. A distinction can
be made here between expert-based and
model-based approaches.

25

26

Assessment and reporting on soil erosion

2.7.1. Expert-based methods


An example of an expert-based approach is
the soil erosion risk map of western Europe
by De Ploey (1989). The map was produced
by various experts who delineated areas
where, according to their judgment, erosion
processes are important. A limitation of this
approach is that the author does not give a
clear-cut definition of the criteria according
to which areas were delineated (Yassoglou et
al., 1998).
Factorial scoring is another approach that
can be used to assess erosion risk (Morgan,
1995). The Corine soil erosion methodology
produced soil erosion risk maps with a
resolution of 1 km for southern Europe
(Corine, 1992), excluding northern Europe.
A relative ranking of soil erosion risk per area
was obtained through the summation of
individual erosion risk scores for each of the
following parameters: rainfall, soil
susceptibility, slope angle, slope distance,
land use and prevention measures. The
Corine approach relies heavily on risk
assessment by experts, and it remains difficult
to assess the effect of changes in land use
and/or climate on the erosion risk as no
quantitative estimate of soil erosion is made.
For the same reasons, it is not feasible to
incorporate more detailed data, nor is it
possible to evaluate the accuracy of the final
result. More details are provided in Section
6.3.
Montier et al. (1998) developed an expertbased method for the whole of France. As
with Corine, the method is based on scores
that are assigned to factors related to land
cover (nine classes), the soils susceptibility to
surface crusting (four classes), slope angle
(eight classes) and erodibility (three classes).
An interesting feature of their method is that
it takes into account the different types of
erosion that occur on cultivated areas,
vineyards, mountainous areas and the
Mediterranean. This way, the interaction
between soil, vegetation, slope and climate is
accounted for to some extent.
A problem with most methods based on
scoring is that the results are affected by the
way the scores are defined. In addition to
this, classifying the source data in, for
example, slope classes results in information
loss, and the results of the analyses may
depend strongly on the class limits and the
number of classes used. Moreover, unless
some kind of weighting is used each factor is
given equal weight, which is not realistic. If

one decides to use some weighting, choosing


realistic values for the weights may be
difficult. The way in which the various factors
are combined into classes that are functional
with respect to erosion risk (addition,
multiplication) may also pose problems
(Morgan, 1995). Finally, as factorial scoring
produces qualitative erosion classes, the
interpretation of these classes can be
difficult.
2.7.2. Model-based methods
A wide variety of models are available for
assessing soil erosion risk. Erosion models
can be classified in a number of ways. One
may make a subdivision based on the
timescale for which a model can be used:
some models are designed to predict longterm annual soil losses, while others predict
single storm losses (event-based).
Alternatively, a distinction can be made
between lumped models that predict erosion
at a single point, and spatially distributed
models. Another useful division is the one
between empirical and physical-based
models. The choice for a particular model
largely depends on the purpose for which it is
intended and the available data, time and
money.
Jger (1994) used the empirical universal soil
loss equation (USLE) to assess soil erosion
risk in Baden-Wrttemberg (Germany). De
Jong (1994) used the Morgan, Morgan and
Finney model (Morgan et al., 1984) as a basis
for his Semmed model. Input variables are
derived from standard meteorological data,
soil maps, multi-temporal satellite imagery,
digital elevation models and a limited
amount of field data. This way, erosion risk
can be assessed over large, spatially diverse
areas without the need for extensive field
surveys. So far, the Semmed model has been
used to produce regional erosion risk maps
of parts of the Ardche region and the Peyne
catchment in southern France (De Jong,
1994; De Jong et al., 1998).
Kirkby and King (1998) assessed soil erosion
risk for the whole of France using a modelbased approach. Their model provides a
simplified representation of erosion in an
individual storm. The model contains terms
for soil erodibility, topography and climate.
All storm rainfall above a critical threshold
(whose value depends on soil properties and
land cover) is assumed to contribute to
runoff, and erosion is assumed to be
proportional to runoff. Monthly and annual

Part I Assessment and reporting on soil erosion

erosion estimates are obtained by integrating


over the frequency distribution of rainstorms.
Several problems arise when applying
quantitative models at regional or smaller
scale. First, most erosion models were
developed on a plot or field scale, which
means that they are designed to provide
point estimates of soil loss. When these
models are applied over large areas the
model output has to be interpreted carefully.
One cannot expect that a model that was
designed to predict soil loss on a single
agricultural field produces accurate erosion
estimates when applied to the regional scale
on a grid of say 50 km pixels or coarser. One
should also be aware of which processes are
actually being modelled. For example, the
well-known universal soil loss equation
(USLE) was developed to predict rill- and
interrill erosion only. Therefore, one cannot
expect this model to perform well in areas
where gully erosion is the dominant erosion
type, let alone mass movements like
landslides and rockfalls.
Also, at the regional scale it is usually
impossible to determine the models input
data (like soil and vegetation parameters)
directly in the field. Usually, the model
parameters are approximated by assigning
values to mapping units on a soil or
vegetation map, or through regression
equations between, for example, vegetation
cover and some satellite-derived spectral
index. In general, however, this will yield
parameter values that are far less accurate
than the results of a field survey. Because of
all this, the relative soil loss values produced
by models at this scale are generally more
reliable than the absolute values.
This is not necessarily a problem, as long as
the user is aware that the model results give a
broad overview of the general pattern of the
relative differences, rather than providing
accurate absolute erosion rates. Because of
this, the availability of input data is probably
the most important consideration when
selecting an erosion model at the regional/
national scale. It would not make sense to use

a sophisticated model if sufficient input data


are not available. In the latter case, the only
way to run the model would be to assume
certain variables and model parameters to be
constant. However, the results would
probably be less reliable than the results that
would have been obtained with a simpler
model that requires less input data (De Roo,
1993). Also, uncertainties in the models
input propagate throughout the model, so
one should be careful not to use an overparameterised model when the quality of the
input data is poor.
Perhaps the biggest problem with erosion
modelling is the difficulty of validating the
estimates produced. At the regional and
larger scale, virtually no reliable data exist for
comparing estimates with actual soil losses.
King et al. (1999) attempted to validate an
erosion risk assessment for France by
correlating soil loss with the occurrence of
mudflows. However, other processes are
involved here and such comparisons do not
substitute for real measurements.

2.8. General conclusions of review


The proposed indicators for soil erosion are
evaluated according to the OECD criteria
listed in Section 2.1. The six pressure
indicators, average farm size, average field
size, consumption of fertiliser, number of
grazing animals, crop yield and net profit,
relate to agricultural intensification. These
pressure indicators all have in common that
they are complex and not directly linked to
the phenomenon of soil erosion. The
identified indicators of state and impact are
difficult or expensive to measure and the
data are usually not readily available.
Indicators of response are prevention and
control measures, which are rarely in
existence at present or are not recorded.
One major conclusion is that the intensity of
agriculture should never be evaluated alone
in relation to erosion. Soil loss due to erosion
is a result of climate, topography, soil
properties, land cover and land
management. Land cover also includes the
natural vegetation.

27

28

Assessment and reporting on soil erosion

3. Driving force, pressure and state


indicators related to land use
Accelerated soil erosion, in excess of natural
geological rates, is caused by anthropogenic
activity. Human activity is a major factor in
shaping the landscape, whereas the physical
structure of a landscape often constrains its
use. Land use and management are the result
of these human activities and as such are the
most important factors that influence soil
erosion. This chapter focuses on the link
between soil erosion and land use, and how
this link is and/or should be reflected in the
proposed indicators.

3.1. Soil erosion indicators and land


use
Driving forces related to soil erosion have
been defined on the basis of intensification
of agriculture in which the risk for
insufficient sustainable land use practices
increases. The proposed indicators aim to
describe the intensification of agriculture
and its increase. As indicators for driving
forces/pressures related to the intensity of
agriculture have been proposed: (a)
consumption of fertilisers per region, (b)
average farm size per region, (c) average
field sizes, (d) average crop yield per area,
(e) average net profit per area, and (f)
number of grazing animals. The intensity of
agricultural production or the trend toward
intensification has been considered as one of
the main causes for soil degradation by
erosion. Furthermore, tourism has been
considered as an important driving force
causing pressures on soil resources. Tourism
has direct impacts in both soil sealing and
soil erosion.
State indicators proposed by the EEA aim to
provide information on the extent of the
area affected by soil erosion. These indicators
represent the short-term approach to soil
erosion assessment. Data for the short-term
approach are mainly derived from
questionnaires from statistical institutions
(Eurostat) and the EEA. Since in most
European countries there are no data
available on soil erosion, such indicators have
limited application.
As an alternative, indicators providing
information on the area under potential and

actual soil erosion risk are proposed. Risk


indicators aim to provide a picture of the
extent and the severity of the potential soil
erosion risk (taking into account climatic,
topographic and soil conditions) and the
actual soil erosion risk (taking into account
the vegetation cover and the actual land use).
The comparison of the potential with the
actual soil erosion risk is a measure for the
impact of land use changes on soil erosion
risk. The information can cover the area of a
member country or the total area in Europe
and provides estimates of the area at risk.
EEA-ETC/S proposed that the long-term
approach to soil erosion assessment should
take into consideration the ground cover due
to vegetation and other protection measures
(e.g. mulching), in areas of a high potential
soil erosion risk.
The long-term indicator approach for soil
erosion is based on data used in the universal
soil loss equation (USLE), such as climate,
soil, relief, vegetation and protection
measures. Problems of such a methodology
are the high variability in space and time of
data such as ground cover, type of land use,
and protection measures.
The Corine soil erosion project has identified
land use and vegetation cover as a major
input in defining actual soil erosion risk.
Land use and vegetation maps are normally
highly generalised from one area to another,
and are quickly out of date. Remote sensing
products provide temporal information and
are proposed here as an important source of
information for vegetation cover. The
normalised difference vegetation index
(NVDI) is often used as an indicator of
vegetation growth determined by optical
sensors. This index when compared during
different periods of the year can indicate the
vegetation cover change during the growing
period of crops and natural vegetation.
However, the introduction of the NDVI will
only make sense if it is combined with
regularly updated land use data, such as
established in the Corine land use map.

Part I Assessment and reporting on soil erosion

3.2. Review of the proposed


indicators in relation to land use
intensity
The proposed indicators in relation to land
use and intensity of land use partially satisfy
the needs for assessing the soil erosion risk
across different agro-ecological regions. As
mentioned in the EEA-ETC/S working
report (EEA-ETC/S, 1999), data on soil
erosion made available by the EU countries
are highly variable. Furthermore, the
application of the universal soil loss equation
has the disadvantage of requiring data such
as vegetation cover at a high temporal and
spatial resolution.
The proposed indicators of intensification of
agriculture can be considered as a good basis
for assessing soil erosion risk but they require
further expansion with other factors such as
other human activities that affect land cover,
existing policies for the protection of soils
and the degree of enforcement of such
policies.
It has to be considered that in hilly cultivated
areas tillage erosion is usually much more
important than wind and water erosion. In
the last decades there is an increasing
awareness that the erosion processes which
are primarily responsible for the severe
degradation occurring in topographically
complex landscapes cannot be attributed to
wind or water erosion only, but are caused
mainly by tillage erosion. Tillage erosion is a
progressively downslope translocation of soil
caused mechanically by tillage implements,
and it is considered as a main cause of land
degradation and land abandonment in hilly
cultivated areas throughout the EU
countries. Areas that have been introduced
to cultivation during this century are being
abandoned at an increasing rate in the last
decades due to a dramatic decrease of the
land productivity resulting mainly from
tillage erosion. The availability of heavy
powerful machinery has favoured deep soil
ploughing with high speeds, and in
directions usually perpendicular to the
contour lines, causing displacement of huge
amounts of soil from upper landscape
positions and deposition to lower landscape
positions. Tillage erosion exposes subsoil,
which may be highly erodible by wind or
water, and fills in ephemeral flow areas,
acting as a delivery mechanism for water
erosion. Data from various sources show that
tillage erosion can account for up to 70 % of

the total loss in cultivated areas (Van Muysen


et al., 1999).

3.3. Options for the future on


relating land use and land use
intensity to soil erosion
The rate of soil degradation is dependent
upon the rate of land cover degradation,
which in turn is influenced by both adverse
climatic conditions and land use
management changes. Vegetation cover, type
of land use, and intensity of land use are
clearly important factors controlling the
intensity and the frequency of overland flow
and surface wash erosion. Vegetation cover
may be altered radically by man within a
short time, but physical and biological
changes within the soil, affecting erosion
rates, may take longer periods. Type of land
use and land use intensity is affected by
various environmental and socioeconomic
factors, therefore indicators for soil erosion
risk assessment should be related to these
factors.
3.3.1. Climate characteristics affecting
vegetation
The characteristics of the climate of an area
that can affect vegetation growth and
vegetation cover and therefore soil erosion
are rainfall, both amount and intensity, and
aridity. These climate characteristics are
easily available for all regions of the EU.
Erosion data collected in various sites along
the Mediterranean region show that the
amount of rainfall has a crucial effect on soil
erosion. Generally, there is a tendency of
increasing runoff and sediment loss with
decreasing rainfall in hilly Mediterranean
shrublands, especially in the region where
rainfall is greater than 300 mm/year. Below
the 300 mm annual rainfall limit, runoff and
sediment loss decrease with decreasing
rainfall. Rainfall amount and distribution are
the major determinants of biomass
production on hilly lands. Decreasing
amounts of rainfall combined with high rates
of evapotranspiration drastically reduce the
soil moisture content available for plant
growth. In areas with annual precipitation of
less than 300 mm and high
evapotranspiration rates, the soil water
available to the plants is reduced drastically.
The soil remains relatively bare favouring
overland water flow whenever rainfall events
happen.

29

30

Assessment and reporting on soil erosion

Aridity is a critical environmental factor in


determining the evolution of natural
vegetation by considering the water stress,
which may occur and cause reduced
vegetation cover. In the Mediterranean
region, vegetation presents a great capacity of
adaptation and resistance to dry conditions,
and many species can survive many months
through prolonged droughts with soil
moisture content below the theoretical
wilting point. Aridity can greatly affect plant
growth and vegetation cover, particularly
annual plants. Under dry climatic conditions
in areas cultivated with rain-fed cereals, the
soil remains bare favouring high erosion
rates under heavy rainfalls following a long
dry period.
Closely related to climatic characteristics is
the topographic attribute, slope aspect. Slope
aspect is considered an important factor for
land degradation processes. Aspect affects
the microclimate by regulating the angle and
the duration at which sunrays strike the
surface of the soil. In the Mediterranean
region slopes with southern and western
facing aspects are warmer, and have higher
evaporation rates and lower water storage
capacity than northern and eastern aspects.
Therefore, a slower recovery of vegetation
and higher erosion rates are expected in
southern and western aspects than in
northern and eastern aspects. As a
consequence, southern exposed slopes
usually have a persistently lower vegetation
cover than northern exposed slopes. The
degree of erosion measured along southfacing hill slopes is usually much higher
(even twice higher) than in the north-facing
slopes under various types of vegetation
cover.
3.3.2. Vegetation characteristics affecting soil
erosion
Indicators of soil erosion related to the
existing vegetation can be considered in
relation to: (a) fire risk and ability to recover,
(b) erosion protection offered to the soil,
and (c) percentage plant cover. Forest fires
are one of the most important causes of land
degradation in hilly areas of the
Mediterranean region. Fires have become
very frequent especially in the pinedominated forests during the last decades
with dramatic consequences in soil erosion
rates and biodiversity losses. The frequency
of fire occurrence is lower in grasslands and
mixed Mediterranean macchia with
evergreen forests. Also, Mediterranean
pastures are frequently subjected to man-

induced fires in order to renew the biomass


production. The Mediterranean vegetation
type is highly inflammable and combustible
due to the existence of species with a high
content of resins or essential oils. Conversely,
it is known that vegetation has a high ability
to recover after fire and the environmental
problems related to fire normally last for only
a limited number of years after the fire
occurred.
There are several factors that affect the
process of the recovery, apart from the fire
and site characteristics, which can be both
natural and anthropogenic. Years of unusual
drought or sites that cannot be affected from
the moist sea winds during summer show a
slower rate of recovery. Human interference,
such as livestock grazing or change in the
land use pattern, may damage irreversibly the
recovering vegetation. Particularly important
are the time intervals between subsequent
fires. The ability of the ecosystems to recover
is not unlimited and a fire frequency beyond
a certain threshold can also lead to a
permanently degraded state. This can be due
both to the nutrient and seed bank depletion
and to increased erosion. These processes
have already led to severe degradation of
extensive hilly areas in the Mediterranean
region.
Vegetation and land use are clearly important
factors controlling the intensity and the
frequency of overland flow and surface wash
erosion. Extensive areas cultivated with rainfed crops such as cereals, vines, almonds and
olives are mainly confined to hilly lands with
shallow soils which are very sensitive to
erosion. These areas become vulnerable to
soil erosion because of the decreased
protection by vegetation cover in reducing
effective rainfall intensity at the ground
surface. Almonds and vines require frequent
removal of perennial vegetation using
herbicides or by tillage. In fact, soils under
these crops remain almost bare during the
whole year, creating favourable conditions
for overland flow and soil erosion.
Erosion data measured along the northern
Mediterranean region and the Atlantic
coastline located in Portugal, Spain, France,
Italy and Greece in a variety of landscapes
and under a number of land uses
representative of the Mediterranean region
(rain-fed cereals, vines, olives, Eucalyptus
plantation, shrubland) showed that the
greatest rates of runoff and sediment loss
were measured in hilly areas under vines.

Part I Assessment and reporting on soil erosion

Areas cultivated with wheat are sensitive to


erosion, especially during winter, generating
intermediate amounts of runoff and
sediment loss especially under rainfalls
higher than 380 mm per year. Olives grown
under semi-natural conditions, particularly
where there is an understorey of annual
plants greatly restrict soil loss to negligible
values. Erosion in shrublands increased with
decreasing annual rainfall to values in the
range of 280300 mm, and then decreased as
rainfall decreased further.
Several hilly areas under natural forests
around the Mediterranean region have been
reforested with exotic species such as
Eucalyptus. Such soils are undergoing intense
erosion as compared with soils left under
natural vegetation. However, the measured
rates of erosion under Eucalyptus are
relatively lower than those measured under
vines, almonds and cereals.
Soil erosion data measured from various
types of vegetation and certain physiographic
conditions showed that the best protection
from erosion was measured in areas with a
dominant vegetation of evergreen oaks, pines
and olive trees under semi-natural condition.
Pines have a lower ability to protect the soils
in southern aspects due to the higher rate of
litter decomposition and the restricted
growth of understorey vegetation. Deciduous
oak trees offered relatively low protection
from erosion in cases where the falling leaves
did not cover the whole soil surface.
The main factors affecting the evolution of
the Mediterranean vegetation, in the long
term, are related to the irregular and often
inadequate supply of water, the long length
of the dry season, and in some cases fire and
overgrazing. According to the types of leaf
generation, the following two major groups
of vegetation can be distinguished: (a)
deciduous: drought avoiding with a large
photosynthetic capacity but no resistance to
desiccation; and (b) evergreen
(sclerophyllous): drought enduring with low
rates of photosynthesis. The main response
of the plants to increased aridity is the
reduction in leaf area index. Severe droughts
that cause a reduction in leaf area index may
be beneficial in the short term as plant
respiration is reduced, but such drought will
increase the probability of enhanced soil
erosion when rain eventually falls, as
protective vegetation cover is reduced.

The various ecosystems present in the


Mediterranean region have a great capacity
of adaptation and resistance to aridity, which
most of the species, existing under
Mediterranean climatic conditions, have to
survive. Plants may have to endure soil
moisture contents below the theoretical
wilting point for many months. Most
probably the expected changes in the
vegetation performance, resulting from a
gradual precipitation decrease, would only
be noticed after a critical minimum number
of years.
Among the prevailing perennial agricultural
crops in the Mediterranean, olive trees
present a particularly high adaptation and
resistance to long-term droughts and support
a remarkable diversity of flora and fauna in
the undergrowth. This undergrowth is even
higher than for some natural ecosystems.
Under these conditions, annual vegetation
and plant residues form a satisfactory soil
surface cover, preventing surface sealing and
minimising the velocity of the overland water.
In the case where the land is intensively
cultivated, higher erosion rates are expected.
Many studies have shown that the variation in
runoff and sediment yields in drainage basins
can be attributed to the vegetation cover and
land use management changes. Many
authors have demonstrated that in a wide
range of environments both runoff and
sediment loss decrease exponentially as the
percentage of vegetation cover increases. A
value of 40 % vegetative cover is considered
critical below which accelerated erosion
dominates in a sloping landscape. This
threshold may be modified for different types
of vegetation, rain intensity and land
attributes. It shows, however, that
degradation begins only when a substantial
portion of the lands surface is denuded;
then it proceeds with an accelerated mode
that cannot be arrested by land resistance
alone. Deep soils on unconsolidated parent
materials show slow rates of degradation and
loss of their biomass production potential. In
contrast, shallow soils with lithic contact on
steep slopes have low productivity, and low
erosion tolerance if they are not protected by
vegetation.
3.3.3. Management quality and humaninduced factors
The definition of soil erosion risk of an area
requires both key indicators related to the
physical environment and to the humaninduced stress. A piece of land, irrespective

31

32

Assessment and reporting on soil erosion

of its size, is characterised by a particular use.


This use is associated with a given type of
management, which is dictated mainly by
climate and changes under the influence of
environmental, social, economic,
technological and political factors.
Depending on the particular type of
management, land resources are subject to a
given degree of stress. Moreover, the
existence of environmental policies, which
apply to a certain area, moderates the
anticipated impacts of a given land use type
compared to the situation where no such
policies are in effect.
The extensive deforestation of hilly areas and
intensive cultivation with rain-fed cereals has
already led to accelerated erosion and
degradation in the last century. The erosion
risk is especially high in areas cultivated with
rain-fed cereals. For one or two months after
sowing winter cereals the land remains
almost bare, and the erosion risk is high
considering that rains of high intensity and
occasionally long duration usually occur
during that period. The sloping lands of the
Thessaly plain, the largest lowland of Greece,
were for centuries under grazing especially in
winter by migratory flocks and herds. The
rapid increase in population due to
immigration in the early 1920s resulted in
the sharp increase of the areas, which were
brought under wheat cultivation. Erosion
experiments and estimations from the
exposure of tree roots demonstrated that
erosion on these areas had proceeded at
rates of 1.21.7 cm soil per year since the
introduction of wheat.
Many hilly areas have experienced
abandonment at an increasing rate due to
low productivity. Land abandonment may
lead to the deterioration or replenishment
phase of soils, depending on the particular
land and climatic conditions of the area.
Hilly areas that can support sufficient plant
cover may improve with time by
accumulating organic materials, increasing
floral and faunal activity, improving soil
structure, increasing in infiltration capacity
and, therefore, causing a decrease in the
erosion potential. In cases of poor plant
cover, the erosion processes may be very
active and the degeneration of these lands
may be irreversible. In cases of land partially
covered by annual or perennial vegetation,
the remaining bare land with soils of low
permeability (clays) creates favourable
conditions for overland flow, soil erosion and
land degradation.

In the last few decades, favourable soil and


climatic conditions and the availability of
ground or surface water have resulted in
intensive farming of the lowlands of the
Mediterranean region. The development of
high input agriculture in the plains provided
much higher net outputs than those
obtained from terracing agriculture.
Furthermore, recently the value of such
terraces has markedly declined because of
the low accessibility by tractors. At present,
most of these areas have been abandoned,
and the terraces have collapsed causing a
rapid removal of the soil by runoff water,
apart from where the stonewalls are
protected by the roots of fast-growing shrubs
and trees. Maintaining such terraces appears
a very expensive practice compared to most
other alternatives for soil erosion control.
Wheat production in hilly Mediterranean
areas has drastically declined during the last
few decades and the intensity of grazing has
increased at the same time. Shepherds often
damage the natural vegetation by
deliberately setting fires to eradicate the
vegetation and encourage the growth of new
grass, which the livestock then overgraze.
Once the land is bare of its vegetative cover
and the soil is loosened, the torrential rains
of autumn and winter begin to wash away the
topsoil.
The process of land degradation can be
greatly accelerated by high densities of
livestock which lead to vegetation
degradation and, in turn, to soil compaction.
An obvious consequence of overgrazing is the
increase in soil erosion, since the gradual
denudation of the landscape exposes the soil
to water and wind erosion. Under such
management conditions and hot and dry
climatic conditions, soils of these areas
cannot economically support a sufficient
vegetative cover to avoid degradation.
Overgrazing of these climatically and
topographically marginal areas,
accompanied by fires, constitutes a
degradation-promoting land use, further
depleting the existing land resources.
The recent number and the extent of forest
fires occurring in the Mediterranean region
are amongst the most serious environmental
problems. In addition to the loss of
vegetation, forest fires induce changes in
physico-chemical properties of soils, such as
water repellence, loss in nutrients and
increased runoff and erosion. They also
destroy wildlife habitat, cause loss of human

Part I Assessment and reporting on soil erosion

life and damage infrastructure. The loss of


vegetation after fire and the progressive
inability of soils to regenerate adequate
vegetation cover due to erosion have already
led to severe degradation of hilly areas in the
Mediterranean region.
Fires have become frequent in pinedominated forests during the last 50 years.
Most of the fires can be attributed to the
carelessness of people. The majority of fires
occur in areas with high xerothermic indices
and moisture deficits. Soil dryness and wind
speed are the principal factors of fire
evolution. The areas affected by forest fires
are increasing dramatically throughout the
Mediterranean basin. In the period from
1960 to 1975, the average rate of burning was
200 000 ha/year, from 1975 to 1980 470 000
ha/year, and 660 000 ha/year from 1981 to
1985.
Erosion rates seem to be enhanced after fires.
The increased erosion rates are only partly
due to the removal of vegetation. More
important seems to be the forming of an
impermeable subsurface layer, which
decreases infiltration rates, while causing a
quick saturation of the upper layers leading
to overland flow and erosion. In contrast
aggregate stability increases after fire and
that increase is more pronounced after
severe burns.
The management quality can be related to
the intensity of land use and to the applied
measurements for environmental protection
related to certain policies. Land use can be
classified according to several criteria leading
to hierarchies of land use types. The number
of criteria employed is dictated by the level of
detail desired as well as by the availability of
the proper data. The principal classification
criterion is the main purpose for which land
is used. Based on this criterion, the land use
types can be distinguished as following:
agricultural land (cropland, pasture or
rangeland),
natural areas (forests, shrubland, bare
land),
mining land (quarries, mines, etc.),
recreation areas (parks, compact tourism
development, tourist areas, etc.),
infrastructure facilities (roads, dams, etc.).
Using the above classification of land use on
land parcels allows the intensity of land use
and the enforcement of policy on
environmental protection to be assessed. The

intensity of land use of a cropland can be


evaluated on the basis of the frequency of
irrigation, degree of mechanisation of
cultivation, application of fertilisers and
agrochemicals, types of plant varieties used,
etc. In the degree of mechanisation, the
following characteristics should be included:
type of tillage instrument, plough depth,
wheel speed of the tractor, direction of tillage
operation, etc.
The intensity of land use of a pastureland can
be defined by estimating the sustainable
stocking rate (SSR) and the actual stocking
rate (ASR) for the various land parcels under
grazing. The ratio of ASR/SSR can be used to
assess the intensity of land use.
In natural areas such as forests, shrubland,
etc., the intensity of land use can be defined
by assessing the actual (A) and sustainable
yield (A/S). Then the intensity of land use
can be classified based on the ratio A/S.
The intensity of land use for areas with
mining activities can be defined by evaluating
the measurements undertaken for soil
erosion control such as terracing, vegetation
cover, etc. Then the intensity of land use can
be classified based on the evaluated degree of
land protection from erosion.
In areas undergoing active recreational use
such as skiing, motor rallies, etc., the
intensity of land use can be evaluated by
defining the actual and the permitted
number of visitors per year (A/P). Then the
land use intensity can be classified based on
the ratio A/P.
Particular attention must be given to the
policies related to soil protection such as
policies supporting terracing, policies
favouring extensive agriculture, etc. Of
course their effectiveness depends on the
degree to which they are enforced.
Therefore, rating of policies can be based on
the degree to which they are enforced.
Hence, the information must be collected on
the existing policies and their
implementation /enforcement.

3.4. Conclusions of review of


indicators in relation to land use
Many of the soil erosion indicators proposed
by the EEA relate to land use and land use
intensity. Land use and vegetation cover, in
general, are the major input in defining
actual soil erosion risk. It is therefore

33

34

Assessment and reporting on soil erosion

advocated to use regularly updated land


cover data, such as established in the Corine
land use map, in combination with remotely
sensed products such as the normalised
difference vegetation index (NDVI) in order
to capture seasonal variations in land cover.
The proposed indicators of intensification of
agriculture can be considered as a good basis
for assessing soil erosion risk but they require
further expansion with other factors that
affect land cover, existing policies for the
protection of soils and the degree of
enforcement of such policies. Other human

activities that affect land use and determine


land use intensity include infrastructure,
recreation, mining activities or forest
management.
Land cover is affected by different
environmental and socioeconomic factors,
such as precipitation, vegetation type and
management quality, which require
monitoring in order to understand the
complex relationship with soil erosion.
Concerning management, tillage erosion is a
prime example of human-induced erosion.

Part I Assessment and reporting on soil erosion

4. Regional assessment of the


extent of soil erosion by water
A regional soil erosion assessment, providing
an estimate of the area affected by soil
erosion and the expected magnitude, is
needed in order to make objective
comparisons that may provide a basis for
further environmental analysis, economic
statements or policy development. Suitable
assessment methods need to be developed to
this purpose.
This section deals primarily with assessing the
extent of soil erosion by water as this is the
most important form of soil erosion in
Europe. Four alternative methods for
carrying out regional assessment are
compared. The Glasod maps and hot-spot
map can be classified as methods based on
distributed point data, while the RIVM and
Corine maps can be classified as factor- or
indicator-based maps. A description of the
processes of soil erosion, crucial to an
understanding of the following sections, can
be found in Annex V.

4.1. Alternative assessment methods


Assessments of soil erosion on a European
scale are required for a number of reasons:
1. to make objective comparisons of the soil
resource, taking account of past
erosional degradation;
2. to estimate the average rate of erosion to
estimate the rate of loss of soil resource
and its economic cost;
3. to estimate the probability and
distribution of severe erosion events, to
evaluate the implications for loss of
production and off-site deposition;
4. to provide an objective basis for
allocation of resources for remediation,
mitigation or more detailed research and
assessment;
5. to assess the impact on the soil resource
of future climate and/or land use
change, due to global warming, possible
policy changes and economic conditions.
Assessment of soil erosion may be based on a
range of methodologies. Some of these are

based on the collection of distributed field


observations, others on an assessment of
factors, and combinations of factors, which
influence erosion rates, and others primarily
on a modelling approach. All of these
methods require calibration and validation,
although the type of validation needed is
different for each category. There are also
differences in the extent to which the
assessment methods identify past erosion and
an already degraded soil resource, as
opposed to risks of future erosion, under
either present climate and land use, or under
scenarios of global change.
4.1.1. Distributed point data
One important form of erosion assessment is
from direct field observations of erosion
features and soil profile truncation. Erosion
features consist of rills and gullies, some of
these ephemeral, and associated deposition
in swales and small valleys. Soil profiles may
show local loss of upper horizons, or burial
by deposition from up-slope. Deposited
material may provide dateable material,
which can indicate when erosion occurred,
but much of this evidence is cumulative over
the period since cultivation began, or in
some cases over the whole of the Holocene.
Data may be collected from regional experts
in soils or soil erosion. They may also be
collated from field or remote (air photo)
surveys of erosion features. Higher satellite
resolution (e.g. Ikonos) may, in the near
future, also allow this method to be applied
from space platforms. Some quantitative data
are also available from erosion plot sites.
These methods require validation to
standardise differences in the intensity of
study of different areas and in the clarity of
suitable features on different soil types.
There are also differences in methods and
traditions between scientists in different
areas of Europe. On their own these methods
cannot provide a complete picture except for
small sample areas, and require the use of
other methods to interpolate between areas.
The main advantage of distributed
observations of erosion is that data are
unambiguous where they exist, and give a
good indication of the current state of
degradation of the soil resource, and other

35

36

Assessment and reporting on soil erosion

methods lack this certainty. The main


disadvantage of these methods is that they
provide little or no information about when
erosion occurred, unless there are
supporting data on this point. Many areas of
the Mediterranean are thought to have
suffered anthropogenic acceleration of
erosion since early classical times, and many
hills are now denuded of their former natural
soil cover. Although of great historical
interest, this has little bearing on current or
prospective erosion hazards.
4.1.2. Factor or indicator mapping
Since many of the processes and factors
which influence the rate of erosion are well
known, as outlined above, it is possible to
rank individual factors for susceptibility to
erosion, providing a series of erosion
indicators. For example, climatic indices may
be based on the frequency of high intensity
precipitation, and on the extent of aridity or
rainfall seasonality. Soil indicators may reflect
the tendency to crusting and the
experimental erodibility of soil particles or
aggregates. Similar rank indicators may be
developed for parent materials, topographic
gradient and other factors. Clearly a high
susceptibility for all factors indicates a high
erosion risk, and a low susceptibility for all
factors indicates a low erosion risk.
Individual indicators may be mapped
separately, but it is more problematic to
combine the factors into a single scale, by
adding or multiplying suitably weighted
indicators for each individual factor. There
are difficulties both about the individual
weightings and about the assumed linearity
and statistical independence of the separate
factors. The method should therefore be
most effective for identifying the extremes of
high and low erosion, but less satisfactory in
identifying the gradation between the
extremes.
Despite these theoretical limitations, factor
or indicator mapping has the considerable
advantage that it can be widely applied using
data which are available in Europe-wide GIS
for topography and soils at 1 km resolution,
and for climate at 50 km resolution. Kosmas
et al. (1999) provide one example of this
approach, applied on a regional scale to
areas in Greece, Italy and Portugal.
4.1.3. Process modelling
There is a continuous spectrum between
mapping based on ranked indicators and
process models with a more explicit physical

or empirical basis. Nevertheless it is fruitful


to consider, as a third approach towards
Europe-wide soil erosion assessment, the
application of a process model. Although, at
first sight, this approach appears to be the
most generally applicable, there are major
problems of validation, and in particular in
relating coarse scale forecasts to available
erosion rate data, much of which is for small
erosion plots. Many of the most successful
process models require more detailed
distributed parameter and rainfall intensity
data than are currently available on a
European scale, so that they cannot be
applied without radical simplification. One
important aspect of this problem is the need
to develop a model that can be used for
validation at fine scales, and for Europe-wide
forecasting on a coarse scale, so that crossscale reconciliation must be as explicit as
possible. Nevertheless this approach has the
potential to provide a rational physical basis
to combine factors that can be derived from
coarse scale GIS, and that overcome the
difficulties about weighting and intercorrelation that are encountered in purely
factor-based assessments.
Process models have the potential to respond
explicitly and rationally to changes in climate
or land use, and so have great promise for
developing scenarios of change, and what-if
analyses of policy or economic options. Set
against this advantage, process models
generally make no assessment of degradation
up to the present time, and can only
incorporate the impact of past erosion where
this is recorded in other data, such as soil
databases. Models also generally simplify the
set of processes operating, so that they may
not be appropriate under particular local
circumstances. Although the USLE has been
the most widely applied model in Europe
(e.g. Van der Knijff et al., 2000), it is now
widely considered to be conceptually flawed.
Other models are now emerging, based on
runoff thresholds (e.g. Kirkby et al., 2000) or
the MIR (minimum information
requirement) approach (Brazier et al., 2001)
applied to the more complex USDA WEPP
model (Nearing et al., 1989).

4.2. The Corine approach


The Corine programme was established in
1985:
1. to help guide and implement
Community environment policy, and to
help incorporate an environmental

Part I Assessment and reporting on soil erosion

Potential versus actual erosion risk as estimated by the Corine methodology

Figure 4.1
Source: Corine, 1992.

dimension into other policies, by


providing information on priority topics;
2. to help ensure optimum use of financial
and human resources by organising,
influencing and encouraging initiatives
by international organisations, national
governments or regions to obtain
environmental information;
3. to develop the methodological base
needed to obtain environmental data
which are comparable at a community
level.
The Corine soil erosion risk maps (Figure
4.1) are the result of an overlay analysis by a
geographical information system, enabling
the evaluation of the soil erosion risk
category. The main source of information
used was the soil map of the European
Communities (CEC, 1985). Potential soil
erosion risk was defined as the inherent risk
of erosion, irrespective of current land use or
vegetation cover (Corine, 1992). The map of
potential erosion risk (Figure 4.1) therefore
represents the worst possible situation. The
area of land in this region with a high erosion
risk totals 229 000 km2 (about 10 % of the
rural land surface). The largest area is found
in Spain, mainly in the southern and western
parts. In Portugal, areas of high erosion risk
cover almost one third of the country. About

20 % of the land surface in Greece, 10 % in


Italy and 1 % in France is subject to high
erosion risk. The difference between the
areas of potential and actual erosion risk
(compare maps in Figure 4.1) reflects the
protective influence provided by present land
cover, and the dangers inherent in changes
in land use practices.
4.2.1. Methodology
For one of these priority topics, soil erosion,
a new methodology was developed, which
provides a factor-based assessment of risk
(Figure 4.2). It was recognised that there was
no suitable Europe-wide map of erosion, and
that existing maps differed widely in
methodology and scales of assessment. The
methodology used was based on a
simplification of the universal soil loss
equation (USLE), a regression-based model,
for which there is a massive database for US
conditions, but little systematic data for
Europe.
E=KRSPV

(4.1)

Where E is the annual soil loss,


K is soil erodibility,
R is rainfall erosivity,
S is the slope length factor,
P is the crop management practice
factor,
and
V is the vegetation cover factor.

37

38

Assessment and reporting on soil erosion

Figure 4.2

Methodology for Corine soil erosion assessment

Source: Corine, 1992.


Soil texture, ST
0 for bare rock
1 for C, SaC, SiC
2 for SsCL, CL, SiCL,
Lsa
3 for SaL, L, SiL, Si

Erodibility, K
0 for ST.SD.SS = 0
1 for 0 < ST.SD.SS < 3
2 for 3 < ST.SD.SS < 6
3 for ST.SD.SS > 6

Soil depth, SD
1 for > 75 cm
2 for 2575 cm
3 for < 25 cm

Soil stoniness, SS
1 for > 10 %
2 for < 10 %

Fournier index, F
1 for S pi2S / p < 60
2 for 60 < S pi2S / p < 90
3 for 91 < S pi2S / p < 120
4 for S pi2S / p > 120

Erosivity, R
1 for F.B < 4
2 for 4 < F.B < 8
3 for F.B > 8

Potential soil erosion risk, EP


0 for K.R.S = 0
1 for 0 < K.R.S < 5
2 for 5 < K.R.S < 11
3 for K.R.S > 11

Bagnouls-Gaussen
aridity index, B
1 for S (2TI-pi) = 0
2 for 0 < S (2TI-pi) < 50
3 for 50 < S (2TI-pi) < 130
4 for S (2TI-pi) > 130

Slope angle, S
1 for < 5 %
2 for 515 %
3 for 1530 %
4 for > 30 %
Actual soil erosion risk, EA
0 for EP.V = 0
1 for EP.V = 12
2 for EP.V = 34
3 for EP.V = > = 5
Land cover, V
1 for fully protected
2 for not fully protected

The USLE (equation (4.1)) is intended to


provide an estimate of average annual
erosion loss in tonnes per unit area. The
Corine soil erosion methodology is a
considerable simplification of the USLE
(Corine, 1992; Briggs and Giordano, 1995).
Erodibility is estimated from soil texture,
depth and stoniness. Erosivity is estimated
from the Fournier and Bagnouls-Gaussen
climatic indices. Slope gradient is included,

but without a slope length correction, and


vegetation and crop management are
collapsed into two categories of protected
and not fully protected, using data from the
associated Corine land cover database. These
factors are combined to estimate three
categories of potential and actual soil erosion
risk. Potential risk excludes vegetation
factors, and so identifies land at risk, while
actual risk includes the vegetation factor to

Part I Assessment and reporting on soil erosion

indicate whether the potential is being


realised. A map showing the assessment for
southern Europe is provided in Figure 4.1
and the overall scheme is summarised in
Figure 4.2.
4.2.2. Advantages and limitations
The Corine soil erosion assessment has the
great advantage of simplicity, in that it
provides a clear forecast, on an objective
basis, for the whole of the area studied. The
method is based, at least in principle, on a
well-established technology, the universal soil
loss equation, which has been very widely
used, both in America and worldwide. Being
based on a factor method within a 1 km GIS
base, the method can be applied at a
resolution that allows discrimination within
regional areas. The method correctly
identifies the areas of the Mediterranean that
have the highest risk of erosion. As a product
of its time, it has considerable merit, and
could be improved with the more detailed
land cover classification now available,
providing refinement in the USLE land cover
and crop management factors.
However, the USLE, although still widely
used on account of its simple structure, is
now widely regarded as a post-mature
technology, and cannot therefore be
recommended as the best basis for estimation
of erosion risk. Furthermore, mapping of
USLE forecasts on national scales, for
example for Italy (Van der Knijff et al., 2000)
shows wide discrepancies between Corine
and USLE forecasts, so that Corine may not
even correctly represent the USLE factors.
The Corine report concedes (p. 92) that
future development of this work would allow
more sophisticated models of soil erosion to
be used. Particularly on improving the factors
used in the procedure, notably in the
calculation of erosivity and soil erodibility,
and in the classification of land cover
(Corine, 1992). On a qualitative basis,
comparison of the erosion maps of southern
Europe appear to show too great a
dependence on the climatic factors in
determining erosion risk, with relatively less
weight given to important factors of
erodibility and land cover. For use in the
future, the Corine assessment also has the

(6)

limitation that it is restricted to southern


Europe, whereas present needs for erosion
data apply to the whole of the European area.

4.3. The hot-spot approach


An analysis and mapping of soil problem
areas (hot spots) in Europe was published
in the EEA-UNEP joint message on soil (EEA,
2000) (6). This addresses a number of soil
problems, and only soil erosion aspects are
reviewed in this section. The purpose of the
study was to support the joint message on the
need for a pan-European policy on soil,
identifying hot spots of degradation in
Europe and examining environmental
impacts leading to change and particularly
degradation of soil function. The work
involved compilation and analysis of data
available at the EEA, together with additional
data from the scientific literature. These data
were incorporated into a GIS (ArcView) for
manipulation and display.
The hot-spot map aims to present a kind of
spatial indicator that would enable the
identification of priorities of intervention
and the visualisation of data gaps.
4.3.1. Methodology
For soil erosion, it is recognised that, because
of its patchy distribution in time and space,
and the uneven density and quality of local
measurements, a simple mapping of hot
spots is futile. The map produced has been
developed from earlier maps (Favis-Mortlock
and Boardman, 1999; De Ploey, 1989), based
on local empirical data, as opposed to Corine
or other estimates based on erosion models,
which are considered unsuitable for
application at coarse scales (EEA, 2001a) (see
Figure 4.3). In the hot-spot approach, broad
zones are first identified for which the
erosion processes are broadly similar. Hot
spots are then highlighted within each zone
and associated with the best estimates, from
the literature, for rates of erosion in these
hot-spot areas. The intention is to identify
areas of current erosion risk, under present
land use and climate, as opposed to either
evidence of past erosion, or of the potential
for erosion under some hypothetical
conditions.

An EIONET review of the hot-spot analysis and maps was undertaken in 2001. The results are discussed in
EEA, 2002b.

39

40

Assessment and reporting on soil erosion

Figure 4.3

Probable problem areas of soil erosion in Europe

Source: EEA, 2000; EEA,


2001.

In detail, the data provide general or


particular information about water erosion
for approximately 60 sites or small regions
across Europe, with measured erosion rates,
which could be placed on the map at 35 sites.
Measurements are taken from erosion plots,
fields and small catchments. The data are
then grouped into three broad groups, for
eastern Europe, the Loess belt and southern
Europe, which primarily represent different
land use history, parent materials and climate
respectively. The problems associated with
erosion hot spots are identified as primarily
off-site in the short term, with siltation and
pollution by agricultural chemicals. In the
longer term, loss of soil productivity is seen as
increasingly important.
4.3.2. Advantages and limitations
Although there are advantages in
concentrating on measured empirical data
where these are abundant, and interpolation
can be meaningful, the sporadic distribution

and episodic occurrence of soil erosion


makes it very ill-suited to this approach.
There is, however, scope for combining data
from the literature with ongoing
measurements and estimates from some
factorial or modelling approaches as a means
of rational interpolation. In its present form
the most important information contained in
these maps lies in the considerable
experience of their compilers, which it is
hard to document or quantify.
Within the area of overlap with the Corine
map in southern Europe, the hot-spot map
inherits from the De Ploey map a greater
concentration on parent material as a key
factor in localising significant erosion. It is
also clear that sites of high erosion identified
on this map are definitely areas of high
impact, but that there is no reliable way to
extrapolate these local results, even to their
surrounding area.

Part I Assessment and reporting on soil erosion

Water erosion vulnerability for 2050, according to the baseline scenario (7) by RIVM

Figure 4.4

Source: EEA, 1999a.


Water erosion risk
in agricultural
areas, 2050

20

10

10

20

a
Arctic Oce

30

40

60

50

60

1000 km
60

low
high
very high
not applicable

North
Sea

A t
l a
n t
i c

moderate

O c
e a
n

30

50

50

e l
a n n
C h

B la

ck

Se

40
40

30
30

1000 km

30

60

not applicable

A t
l a
n t
i c

increase

10

10

20

20

a
Arctic Oce

30

30

40

60

50

60

decrease
no change

20

10

O c
e a
n

Change in water
erosion risk in
agricultural areas,
1990 2050

North
Sea

50

50

e l
a n n
C h

B la

ck

Se

40
40

a
30

30

10

4.4. The RIVM approach


As part of a major report on strategies for the
European environment, a baseline
assessment of water erosion was prepared for
1990 (RIVM, 1992). This assessment of
current risk was combined with climate and
economic projections within the framework
of the IMAGE 2 model to generate scenario
projections for 2010 and 2050. This approach
has the advantage of making explicit scenario
projections, a feature lacking in other
(7)

20

30

approaches, but is currently only available at


50 km resolution, so that it cannot readily be
interpreted at sub-national scales. This
approach also has the advantage of
combining physical and economic elements
within a single framework. However, the
value of this integration must be judged on
the reliability of all components, of which
only the soil erosion assessment is addressed
here. The results of this assessment are
shown in Figure 4.4.

In the last EEA state of environment report (EEA 1999a) an increase in the risk of water erosion was expected
by the year 2050 in about 80 % of EU agricultural areas, as an impact of climate change. The increase would
mainly affect the areas where soil erosion is currently severe. These results were produced jointly with the
Commission, based on business-as-usual socioeconomic and energy developments which did not assume
that the Kyoto targets would be met (pre-Kyoto EC energy scenarios).

41

42

Assessment and reporting on soil erosion

4.4.1. Methodology
IMAGE 2 is primarily a global model
composed of 13 sub-regions (Alcamo, 1994).
OECD-Europe and eastern Europe are two of
these regions. IMAGE 2 is an integrated
model designed to simulate the dynamics of
the global society-biosphere-climate system. It
consists of three fully linked sub-models:
energyindustry that computes emissions of
greenhouse gases (GHGs) as a function of
energy consumption and industrial
production; terrestrial environment that
simulates changes in global land cover and
the flux from biospheric GHGs into the
atmosphere; and atmosphereocean that
computes average global and regional
temperature and precipitation patterns.
IMAGE 2 is linked to the MIDAS model for
CO2 emissions and energy demand and
supply; to GEM-3 for population and GDP by
country and by sector; and to WorldScan for
EU-15 GDP.
Water erosion represents a module of the
IMAGE model adapted from the water
erosion model of Batjes (1996) on a x
(approximately 50 km) grid. The water
erosion impact module generates a water
erosion risk index based on three main
parameters: terrain erodibility, rainfall
erosivity, and land use pressure.
The methodology is described below and
summarised in Figure 4.5.
1. Terrain erodibility is based on soil type
and landform, which are regarded as
constant parameters. Land form is
classified into general types (flat,
undulated, mountainous, etc.) by using
the difference between minimum and
maximum altitudes for each grid cell,
using the 10 minute grid elevation data
set of the Fleet Numerical Oceanography
Centre (FNOC) which provides 9 points
per 50 km grid cell. Soil type is derived
from the FAO Soil map of the world and
is composed of soil depth, soil texture,
and bulk density. General averages for
these characteristics are supplied by the
WISE soil profile data set.
2. Rainfall erosivity is represented by the
month with the maximum rainfall per
rain-day. This is considered to be
indicative of rainfall erosion potential.
Data on precipitation and number of wet
days are derived from the IIASA climate
database for mean monthly measured
climate variables from an array of

weather stations for the period 1931 to


1960. Precipitation is considered a
dynamic variable, while the number of
wet days is assumed to remain constant.
3. The potential erosion risk derived from
these two factors is then converted to
actual erosion risk by a land cover factor,
representing the degree of protection
afforded by various land covers
(agricultural crops) from land cover
maps. Natural vegetation with a closed
canopy (e.g. forests) is assumed to
provide optimal protection (no risk) and
natural vegetation with a more open
structure (e.g. shrubs) is assumed to
provide sub-optimal protection (low
risk). Land cover maps for the IMAGE
model are derived from several sources
including Olsons land cover database
and statistical information from FAO.
4.4.2. Advantages and limitations
The main advantage of the RIVM approach
lies in its potential for integration with other
environmental factors within an integrated
model of the physical and economic
environments, and the IMAGE model used is
not evaluated here. Nevertheless these
advantages cannot be fully realised unless the
underlying model modules are themselves of
an acceptable standard.
The RIVM soil erosion model is a factor
model, like Corine, but, although initiated
six to eight years later, is in many ways a still
more simplified approximation to the
imperfect USLE model. If Figure 4.5 is
compared with Figure 4.2, the similarities
and differences are immediately evident. It
may be seen that the soil erodibility takes a
similar form to Corine or USLE, with
components for soil type, and a simplified
gradient and index. The rainfall erosivity
component is seen as an inadequate
representation, which contains neither the
theoretical basis underlying USLE nor the
fair empirical alternatives provided in
Corine. Only land use provides an
improvement on Corine, due to the
availability of better land cover data than
were available early in the Corine project.
The RIVM method exploits the potential,
inherent in any physically based or factorbased assessment, of providing scenario
analysis, through the inclusion of two
dynamic components, the monthly rainfall
totals (affecting erosivity) and land cover
(affecting the assessed actual erosion risk).

Part I Assessment and reporting on soil erosion

Summary of RIVM methodology for water erosion assessment

Land form (from 10 DEM)


Relief range (for 9 points)

Soil type (from FAO map)


Depth
Texture
Bulk density

Potential erosion risk

Rainfall erosivity (IIASA)


Max. rain (per rain-day)

Land cover (IMAGE)


Crop protection
Low for open shrubland
0 for closed canopy forest

The RIVM approach is therefore seen to


share some of the advantages of all methods
which use distributed data sources, by
providing an objective assessment across the
European area. However, neither the 50 km
resolution nor the implementation of the
factors contributing to erosion are seen as
providing a state-of-the-art assessment.

4.5. The Glasod approach


The main objective of the Glasod project was
to strengthen the awareness of decisionmakers on the risks resulting from
inappropriate land and soil management to
the global well-being. To achieve this, the
United Nations environment programme
(UNEP) commissioned the International Soil
Reference and Information Centre (ISRIC)
in 1988 to coordinate a worldwide
programme in cooperation with a large
number of soil scientists throughout the
world to produce, on the basis of incomplete
existing knowledge, a scientifically credible
global assessment of the status of humaninduced soil degradation within the shortest
possible time frame. The task was
subcontracted to correlators in 21 regions to
prepare, in close cooperation with national
soil scientists, regional soil degradation status
maps. These regional maps were correlated
to provide the Glasod world map of soil
degradation.

Actual erosion risk

It is important to recognise the limited aims


of the project, and to observe that Glasod is
the only approach which has, to date, been
applied on a worldwide scale. It is based on
responses to a questionnaire sent to
recognised experts in all countries (Oldeman
et al., 1991). It thus shares with the hot-spot
approach dependence on a set of expert
judgments, but can provide very little control
or objectivity in comparing the standards
applied by different experts for different
areas.
The information and data on erosion and
physical degradation in the Dobris
assessment (EEA, 1995) are based on an
updated version of the European part of the
global assessment of soil degradation
(Glasod) map. For this update (Van Lynden,
1995), questionnaires were sent to scientific
teams in each European country for
comments and additions on the Glasod map.
Not all countries completed and returned
the questionnaires and the degree of detail
of the information received varies greatly. It
must also be noted that the scale of the maps
(1: 10 000 000) limits the detail that can be
shown, providing a minimum resolution of
approximately 10 km. The results are shown
in Figure 4.6.
The Glasod map identifies areas with a
subjectively similar severity of erosion,
irrespective of the conditions which

43

Figure 4.5

Assessment and reporting on soil erosion

40

30

20

10

10

Arc t ic

20

30

Ocean

40

50

Bare

nts

Se

60

70

80

60

rwe

gian

Wh

No

ea

60

ic S

ni
Gulf o
fB
o

ge

Se
a

S ka

d
inlan
of F

50

ti

Irish
Sea

Nort
h
Sea

gat
Katte

Firth
Fort of
h

lf
Gu

k
rr a

y Fir
th

Celt

th

a
c e

Mora

50

it e

Source: EEA, 1998


modified from Van
Lynden, 1994.

Water erosion of soils in Europe according to the Glasod approach

t l
a n
t i
c

Figure 4.6

Ba

The
Wash

ea

sh Channel
Engli Manche)
(La

44

Ba
y
Bis of
cay

Sea of
Azov

as

ia

1:30 000 000

ri

B lack
at

ic

Se

Sea

40

Sea of
Marmara

Ty rr h en i a n
Sea

Water erosion
Loss
of topsoil

Ligurian
Sea

Gulf o
Lions f

Se

40

Aegean
Sea

Ionian
Sea

Terrain
deformation

Sea of Crete

severity extreme
severity strong
severity moderate
severity light

M e d i t e r r a n e a n

S e a

30

not applicable

produced this erosion. For water erosion,


areas are grouped together primarily on the
basis of the severity of topsoil loss. It is clear
from comparison with other maps that there
are substantial differences between the
objective standards applied in different
regions, although parts of southern Spain,
Sicily and Sardinia are described as areas of
high erosion risk in all assessments.

20

30

4.5.1. Methodology
The stages in the production of the Glasod
global map were as follows (Van Lynden,
personal communication, 2001).
In close collaboration with ISSS, Staring
Centre, FAO and ITC, 300 soil scientists
worldwide were contacted and correlators
for 21 designated regions identified.
These collaborators were provided with
guidelines for the assessment of the status
of human-induced soil degradation (1988)
and with a base map (Mercator project;1:10
million average) with loosely defined
physiographic units (polygons).
The assessment consisted of an expert
judgment (following the general guidelines

Nile

10

40

prepared) of degradation status (type,


extent, degree, rate and cause) for
individual polygons on a national/subnational level.
This information was compiled by the
regional correlators.
Final map compilation and publication of
the global map at ISRIC/Staring Centre in
1990.
The resulting map was digitised afterwards
with an attribute database and
supplementary statistics on the extent and
degree of degradation.
Thematic maps, derived from the Glasod
database, were prepared by UNEP/GRID
for inclusion in the World Atlas of
Desertification. The Glasod map and
complementary statistics have been used
and cited in numerous scientific journals
and policy documents of the World
Resources Institute, the International Food
Policy Research Institute, the Food and
Agriculture Organisation of the United
Nations, the United Nations environment
programme, and many others.

Part I Assessment and reporting on soil erosion

Detailed information on the global extent of


human-induced soil degradation, derived
from Glasod, is included in Greenland and
Szabolcs (1994). Results indicate that water
erosion is the dominant degradation process
in Europe, and that less than 10 % is
considered to be strongly or severely
degraded.
4.5.2. Advantages and limitations
The Glasod map is still widely used and
quoted, although its authors and critics alike
recognise the need for a more detailed and
more quantitative assessment. Its virtue was
that it was produced quickly in response to a
demand, and was never intended as more
than an interim assessment. Nevertheless, the
impossibility of making truly objective
comparisons between, and often within areas,
is a major difficulty in interpreting the
results. No expert knows all the erosion sites
within his or her own area with equal
confidence, and scales within each area tend
to be from best to worst, without absolute
scales for objective comparison. Some of
these problems are being partially addressed
in new assessments from ISRIC, which make
use of physiographic units defined by the
SOTER methodology, but the whole
questionnaire approach is fundamentally
flawed by a lack of detailed knowledge and
the impossibility of objective comparisons.

to allow effective interpolation between


available sites. Thus methods based on
questionnaire surveys (Glasod) or erosion
measurement sites (hot spots) are likely to be
inadequate on their own. In addition,
differences between expert assessments and
measurement methods reduce the
comparability between the limited data
available.
Methods based on factors or indicators have
the immediate benefit of accessing
distributed data sources that are available on
a European scale in electronic form (GIS).
These include climate data, DEMs and soil
maps. All of the mapping methods appear to
use implicit or explicit reference to at least
some indicators, particularly to soil
classifications, but only Corine makes explicit
use of an adequate range of relevant
indicators. However, Corine is an
implementation which is imperfect for
historical reasons of data availability, of a
model (USLE) which is now no longer
considered as state of the art. For these
reasons, although it perhaps gives the best
indication of the Europe-wide distribution of
soil erosion of the four methods surveyed, it
is now in need of replacement, and appears
not to represent expert opinion of variations
in erosion rate within each national region.

4.7. Options for the future


Given that there are now improved
methodologies, based on more quantitative
analysis of particular problems, such as soil
erosion, it is unquestionably timely to
abandon this approach, whilst not rejecting
the data from local erosion sites to calibrate
more quantitative models. However, it was
the first comprehensive global overview on
soil degradation that created awareness and
highlighted the need for a more objective
approach and for validation. Updates for
specific regions have been made under the
Soveur and ASSOD programmes (see Annex
III, workshop paper by Van Lynden).

4.6. Comparative assessment of the


four methodologies
None of the four approaches reviewed here
achieves state-of-the-art forecasting for soil
erosion risk assessment across Europe (NB:
an interpolation with a colour for a region
where no observation was made is also a
forecast). Because soil erosion events are
associated with the incidence of storms,
which are patchy in both time and space, site
data must be widespread and long-continued

It is clear that the widespread availability of


GIS data for key controlling variables strongly
favours a factorial or modelling base for
future assessments of soil erosion. The
difficulties associated with a modelling
approach should not, however, be
underestimated. It is essential that a suitable
model should:
1. represent the state of the art in current
understanding of soil erosion;
2. combine sufficient simplicity for
application on a European scale with a
proper incorporation of the most
important processes;
3. have the potential for downscaling to
field or plot scales where explicit
validation can be made with field
monitoring data, to make full use of
experimental sites available.
Current thinking on modelling (COST623
2001) recognises the importance of runoff
forecasting as a critical control on erosion

45

46

Assessment and reporting on soil erosion

loss. Simple runoff models are based on a


runoff threshold or infiltration equation
approach, and vary in complexity from the
RDI model (Kirkby et al., 2000) to the USDA
WEPP (Nearing et al., 1989) model. There is
a trade-off between a simple model, which
can be applied across a continuous range of
parameters, for each cell within a European
grid (as in the EC Pesera project), and a
more complex model, applied for a finite
number of parameter steps, the permutations
of which are then repeated at many sites
across a region (the MIR approach proposed
by Brazier et al., 2001). In either case, there is
then the additional need to ensure that there
is adequate investment in validation against
existing field data, although recognising
their variations in quality and methodology.
Present-day soil erosion models have
substantially aided insight into erosion
processes, but are designed to assess soil
erosion risk at small spatial units. In addition,
these localised studies may not be
representative of the continental and
regional scales required by policy-makers to
set up an adequate soil conservation strategy.
Moreover, it is often technically and
financially unfeasible to acquire the
necessary input data to run detailed soil
erosion models for decision-making at
regional, national or pan-national level. For
application on a broad regional scale,
current models are severely limited by their
high data demand and, in many cases, by a
focus on individual events rather than on
long-term averages or cumulative impact.
This prevents the application of the best
American models, such as WEPP (Nearing et
al., 1989) and Kineros (Woolhiser et al.,
1990) or other EU-funded models, such as
the Eurosem, Eurowise and Medalus
(Medrush) models. The Corine and USLEderived models (RUSLE, etc.) are more
appropriate in their data needs, but all are
now recognised as lacking a physical basis
which can be linked, more or less explicitly,
with current concepts and research in soil
erosion, and which offer the possibility of
direct provision of physical and
socioeconomic scenarios.
The fifth framework project Pan-European
soil erosion risk assessment (Pesera) will
produce a regional model with a physical
basis that can be applied to larger areas and
can be used for scenario analysis and impact
assessment. Earth observation techniques
and the increased use of geographic
information systems have greatly improved

the availability and methods to process and


analyse spatial data. In concert with the
improved understanding of soil erosion
processes, the development of a spatially
distributed process-based model to assess soil
erosion risk over large areas is therefore the
next challenge. In the face of an inevitable
uncertainty, the concern will be to safeguard
the models robustness based on a welldeveloped strategy of sensitivity analysis.
Measured soil erosion data will play a crucial
role in evaluating the model through quality
assurance in the absence of any
measurements. The model to be developed
will produce quantitative results with a
known reliability, and can be continuously
upgraded with more accurate or detailed
data upon their availability. The latter will
evoke the somewhat underestimated
challenge of reconciling the model with
high-volume data sets. More details on the
Pesera project can be found in Annex III
(Ann Gobins and Mike Kirkbys
presentation).

4.8. Conclusions and


recommendations on
implementation of regional
assessments
A number of recommendations can be made
from the assessment of existing methods in
this section. The most immediate is that
there is scope and need for an improved
assessment method, since all show serious
shortcomings, and only a moderate level of
agreement about the areas most seriously
affected by soil erosion in Europe. The scope
for a new assessment is based on the
emergence of better models at appropriate
scales, which can build on the data and
expertise developed through the Corine
project, to develop a physically based forecast
for the distribution of water erosion across
Europe. The need for a new assessment is
based on the large variation between current
maps, which show no clear consensus on the
areas most at risk.
Additional recommendations relate to the
specification of erosion risk, which is defined
in significantly different ways for the various
assessment methods. It is suggested that
evidence of historical erosion should be used
to modify soil databases, and as a gross
qualitative indicator that an area is
susceptible to erosion under certain
circumstances (which may no longer apply).
All of this information should be included in
an assessment of the existing soil resource,

Part I Assessment and reporting on soil erosion

and this is considered to be separate from an


assessment of soil erosion risk.
Soil erosion risk refers to the expectation of
future loss, under both present conditions
and under different climate (due to global
change) and land use (due to economic
circumstances, global change or policy
implementation). This can most usefully be

expressed in two ways. First as an estimate of


long-term average rates of soil loss, and
second as the loss expected in an extreme
event (for example with 100 years average
recurrence interval). These assessments can
then be directly related to the long-term loss
of soil resources in relation to present soil
depths, and to the likely costs of locally severe
off-site deposition and pollution.

47

48

Assessment and reporting on soil erosion

Part II Workshop conclusions


The soil erosion workshop was held at the
European Environment Agency
(Copenhagen) on 27 and 28 March 2001. It
was attended by about 15 people: EEA staff,
national experts and representatives of the
Commission (Environment DG, JRC).
The workshop was organised to facilitate the
expert review of EEA work on soil erosion
and to make recommendations for the
further development of the work, in
particular to be carried out by the new ETC
on Terrestrial Environment.
The focus of the workshop was on assessment
and reporting of soil erosion indicators
across Europe. This entailed a review of the
European Environment Agencys work,
which specifically concentrated on soil
erosion indicators, including the hot-spot
maps in comparison with other regional
assessment methods (see Part I).
The EEA presented the objective of the
workshop and the state of play of EEA work
on soil, with particular reference to
indicators on soil erosion. A selected number

of experts presented work related to the


assessment of soil erosion and the
development of indicators (see Annex I). A
short discussion followed after each
presentation.
The presentations were divided under three
major headings: soil erosion indicators and
assessment framework, regional and spatial
assessment methods of soil erosion
(indicators of state), and data availability for
soil erosion indicators. A specific discussion
was held immediately after each session,
guided by the questions formulated in Annex
I. During the general discussions emphasis
was placed on indicators of state and impact
for assessing soil erosion.
This part of the report presents an overview
of workshop discussions and
recommendations, organised by theme and
presentation. A separate section deals with
the general discussions. The programme of
the workshop is included in Annex II. The
papers presented are included in Annex III.

Part II Workshop conclusions

5. Soil erosion indicators and


assessment framework
5.1. Operational framework
As soil erosion has impacts on several media,
in particular on water quality, working links
should be developed with other ETCs and
specifically with the ETC on Water. The
Water Framework Directive recognises the
relevance of agriculture as a major source of
water pollution.
Moreover, it was recommended that working
links with groups of experts contributing to
the development of international initiatives
such as the COST Action 623 Soil erosion
and global change, the European Society for
Soil Conservation and IGBP-GCTE Focus 3
Soil erosion network (COST623 2001)
should be maintained.

5.2. Soil erosion indicator work at


ETC/Soil
The soil erosion indicator work at the former
ETC/Soil is described in detail in the
working report by EEA-ETC/S (1999)
prepared by Dwel and Utermann. For soil
erosion indicator development, use was made
of existing databases, which were augmented
with questionnaire obtained data. The
general conclusion is that data availability is
not a real problem, but the accessibility
certainly is.
There is also a perceived need to provide
details on type and methods of erosion data
collection (metadata). This will help establish
an approach to deciding which indicators of
erosion are needed and how to get them,
since we need indicators which tell us what is
happening now (i.e. actual) and what may be
happening now and in the future (i.e.
potential).
A definition of soil erosion is needed. What
type of erosion is described or measured? Is it
current or past (geological)? Present-day
erosion is what interests the policy-maker.
Sediment delivery ratios or sediment loads in
rivers are not all directly related to soil
erosion (part may be caused by riverbank or
channel erosion). Moreover, there is a huge
time lag between conservation measures and
sediment measurements so that care has to

be taken in the interpretation of sediment


concentrations as an indicator of impact
(EEA-ETC/S, 1999). Sedimentation (and the
link to eutrophication) of lakes is also
important and necessitates a link with water
quality studies.

5.3. GISCO databases and tools to


derive pressure indicators for
soil erosion
The ETC/Soil proposed intensification of
agriculture as a major driving force for soil
erosion (EEA-ETC/S, 1999). Subsequently,
indicators for agricultural intensification
were drawn from available European-wide
databases. The advantages are that data are
readily available and serve the short-term
purpose of monitoring.
An understanding of socioeconomic driving
forces and soil erosion is still very limited.
The links between agricultural intensification
and soil erosion are not always clear.
Moreover, agricultural intensification and
soil conservation are not mutually exclusive.
The mechanisms of intensification are also
poorly understood. For example, in the
Belgian Loess belt, land under steep slopes
has been taken out of production while the
agriculture in general has intensified.
However, in the United Kingdom, the more
intensive kinds of agriculture are generally
driving erosion, such as enlargement of
fields, continuous arable, switch to winter
cereals, more irrigation, more crops grown
under plastic, more grazing animals, and
more pigs reared out of doors. Land is only
taken out of cultivation if farmers are paid to
do that, e.g. set-aside, or if they are paid to
reduce grazing intensities. All the evidence
suggests that there was very little erosion
prior to the late 1940s.
A major critique is that indicators of pressure
related to agriculture should take account of
crop types, cropping calendars and
management and crop growth in a spatial
context. The MARS project uses this type of
information for running the crop growth
monitoring system (CGMS) across Europe.
Changes in areas under the various crop
types are available every two years from the
farm structure survey, but only at NUTS 2

49

50

Assessment and reporting on soil erosion

level. The NUTS 3 level agricultural census


data are available at 10 yearly intervals for the
EU, but individual EU Member States collect
this level of information more regularly.
The knowledge that crop types are changing
can help evaluate whether erosion is likely to
get more widespread, or not. Provided that
within individual countries it is not too
expensive nor too difficult to get data on
cropping the changes in erosion risk can be
assessed and passed on to the EEA. This may
circumvent the lack of European-wide
detailed information on cropping from year
to year.

5.4. Discussion on questions


What is soil erosion? It was recommended to
concentrate on present-day soil erosion for
policy purposes. Because of erosions
patchiness, rates are only meaningful for very
small areas. Policy-makers are interested in a
European-wide assessment of the problem at
present and in the future. This requires a
regional assessment in terms of soil erosion
risk.
Mapping actual erosion will be a very timeconsuming and costly operation. Moreover,
the recognised patchiness in time and space
will always call for continuous updates. A risk
assessment will enable a transparent and
objective comparison between regions. The
underlying model in a risk assessment
translates what experts use into mathematical
algorithms. However, a mapping instruction
to map out actual erosion features on a
detailed scale could be an option where more
details are required on the actual state of the
problem and where funds are available to
undertake this expensive operation.

One very important remark is that a


programme to monitor soil erosion across
different agro-ecological regions and under
different land uses should underpin both
mapping exercises and regional soil erosion
risk assessment methods. Only then a sound
approach is ensured of estimations and
mapping features that are directly validated
and compared with measurements.
Moreover, measuring campaigns may lead to
new insights and therefore to better mapping
and risk assessments.
Indicators of state and impact are the most
important. However, factors underlying the
causes of soil erosion such as pressure
indicators should be clarified and
communicated to the policy-maker. It is
important to link each indicator to the
general policy framework. Headline
indicators and sub-indicators should be
identified and prioritised. A major concern is
the link between different indicators of one
category (e.g. driving forces) that is not
expressed nor explored within the DPSIR
assessment framework.
Agriculture in general is a very important
driving force for soil erosion. An example for
some of the less-favoured areas in the
Mediterranean showed that with increasing
subsidies stocking rates increased and
resulted in overgrazing and subsequently
more erosion. A similar scenario is
foreseeable if farmers are compensated for
soil erosion. In a situation of financial
compensation for soil eroded land,
incentives for farmers to practise soil
conservation will be lacking. Care should be
taken in formulating the necessary remedial
measures and encourage farmers in
practising soil conservation techniques.

Part II Workshop conclusions

6. Regional and spatial assessment


methods of soil erosion and data
availability
6.1. The Glasod map
The Glasod map produced by ISRIC was the
first effort to produce a global assessment of
human-induced soil degradation on the basis
of incomplete expert knowledge within the
shortest possible time frame. This approach
provides an overview on a global scale of
human-induced soil degradation, and can be
used to identify hot spots and awareness
raising for international policy-makers. Major
critiques regarding the map relate to the
methodology and are reflected in a strong
correlation between country boundaries and
erosion risk. The Glasod approach has been
further developed by other programmes such
as ASSOD (Assessment of soil degradation in
south and south-eastern Asia, Van Lynden
and Oldeman, 1997) and Soveur (Assessment
of soil degradation in central and eastern
Europe, Batjes, 2000) that are linked to GIS
and database technologies. But Glasod
reliance on qualitative data means that the
approach should not be adopted in isolation.
Another suggested approach is to use the 1:1
million scale soil map of Europe as a base for
a rapid Glasod-type assessment of erosion, i.e.
what is the type of erosion and its extent
within a soil map unit and what are the
causes of that erosion. A major disadvantage
of expert mapping is that the policy-maker
does not know what the underlying criteria
were to produce the map (did the expert use
soils, land use or a combination, etc.).

EU. Additionally, because of erosions spatial


patchiness, it is problematic to link erosion
rates measured at specific locations with the
severity of erosion in the hot-spot areas.
Moreover, spatial links between the different
hot-spot areas are difficult to establish.
Therefore the usefulness of the hot-spot map
to policy-makers was questioned.
The problem with the hot-spot map is not its
aim, i.e. to bring out where erosion is, or is
most likely to occur, but its scale. Small areas,
e.g. soil landscape units cannot be brought
out at this scale. However, a framework such
as a 1:1 million, or preferably a 1:250 000,
soil/land use map could form the basis for
assessing erosion on which hot spots could be
portrayed, i.e. very often erosion will equate
with particular soil/land use associations.
Such a map could be similar to the actual
and potential erosion maps produced for
England and Wales which classify soil/land
use associations from very low, low, medium,
high to very high risk. It would bring
together both expert views as well as
quantitative work. Such an approach would
bring together both the Glasod and hot-spot
methodologies.
An evaluation of the hot-spot map was
carried out by EIONET in spring 2001. The
results of the evaluation are published in
EEA, 2002b.

6.3. Regional assessment of the


extent of soil erosion by water

6.2. The hot-spot map


The hot-spot map (EEA, 2001a) is an
empirical approach using measured data and
expert opinion. This approach was adopted
in view of the difficulty modelling
approaches have in dealing with erosions
spatial and temporal variability, and the
generally poor job these models make of
modelling gully erosion. Three categories are
presented: zones (expert opinion); hot-spot
areas (based on the De Ploey map); and
locations (published erosion rates). An
obvious disadvantage is that there is a lack of
reliable data to give an adequate picture of
erosion hot-spot locations across the whole

A comparison of existing maps for soil


erosion assessment on a European scale was
made. Four specific approaches Glasod,
hot spots, RIVM and Corine, used by the EEA
to obtain a European-wide assessment of soil
erosion, were related to two assessment
methods (distributed point data and factor
or indicator mapping). Regional process
modelling (RDI model and Pesera model)
was presented as a suitable alternative for
future regional erosion risk assessment. A
description of the RDI and Pesera model is
given in Annex III (regional soil erosion risk
assessment by Anne Gobin and Mike Kirkby).

51

52

Assessment and reporting on soil erosion

It was pointed out that the four approaches


used by the EEA each served specific but
different purposes. For instance, the hot-spot
map was aimed at locating soil erosion
problem areas, whereas the RIVM map
illustrated the impacts of global change on
soil erosion. However, the common objective
of all these specific approaches (including
regional process modelling) is a regional
assessment of soil erosion. Some of them
consider soil erosion risk (RIVM, Corine),
while others consider actual erosion (Glasod,
hot-spot map).
The process modelling method has the
advantage of producing an indicator of state
with the possibility for analysing different
scenarios and assessing impacts (i.e. estimate
what may happen in the future). The major
objective of scenario analysis is to reduce soil
erosion through policy-making. Clearly, the
policy-maker has a major impact on land
cover and land management through various
land use policies. Any changes in these two
factors affect soil erosion. The focus remains
on indicators that are relevant to human
activities.
Modelling efforts should be thoroughly
validated against erosion measurements, and
a clear distinction should be made between
modelled erosion risk and present-day
erosion rates. Rainfall intensity is a crucial
input to any soil erosion model and is
incorporated in the Pesera model through
rain distribution as a surrogate (Kirkby et al.,
2000).
All current erosion modelling approaches
have severe limitations in capturing erosions
spatial variability. Moreover, any model based
on the USLE will not include gully erosion,
which, as Poesen et al. (1996a, 1998) have
shown, can be a major contributor to total
erosion in (at least some parts of) Europe.
The Pesera model was presented as a
potential solution to future regional erosion
risk assessment.
Three different views were adopted among
the workshop participants concerning
regional erosion assessment: (1)
representation of real measurements, (2)
expert judgment and (3) modelling
approaches (whether factorial or physically
based). The first two groups represent antimodelling views. There was a general
consensus, however, that both measurements
and expert judgment remain a vital part in
factorial or process modelling of regional

erosion. It was the role or weight that is given


to the different components in the process of
regional assessments that remained a point of
discussion.

6.4. General discussion on regional/


spatial soil erosion indicators
For policy purposes, there is a need to define
a method which could be used to assess the
present state of soil erosion but also to
predict future responses. This calls for the
definition of an indicator and a calculation
procedure.
The use of expert-based maps versus
indicator or model-based maps was discussed
in detail. Experts could judge the severity of
soil erosion and have considerable
knowledge of the detail. However, an
indicator reflecting the extent (area) affected
by soil erosion requires some kind of
interpolation. Moreover, in expert
judgments, the methodology is seldom
repeatable and hence it does not provide a
sound basis for comparing subsequent future
judgments. Expert judgments could be too
biased for policy decisions to be based on.
On the other hand, for a process as complex
as soil erosion, experts have a crucial role to
play in validation and evaluation of the end
product.
Generally, it was agreed that the indicators to
be developed should have the following
properties and adhere to the following
principles:
1.
2.
3.
4.

quantitative,
objectively calculated,
validated against measurements,
evaluated by experts.

It needs to be clearly pointed out that


validation of the model/indicator is essential,
and that this validation needs to be planned
and resourced in the same way as the model/
indicator development itself. Since only
limited measured data are available for the
EU (and the measurements are of rates on
small areas), databases of measurements
need to be combined with expert opinion in
the model/indicator validation. An input in
validation could include radionuclide
measurements to determine current erosion.
Whenever possible, the use of maps should
be accompanied with an accuracy evaluation.
Well-defined, unambiguous class definition
and the analysis of errors may improve the

Part II Workshop conclusions

statistical use of the maps. The mixture of


land use and land cover, like in the Corine
legend, can determine loss of information
and may pose difficulties in the calculation of
indicators. For example, the urban green
area is a category of land use and for the
cover it can be a lawn, a wood area and part
of it could be covered by artificial
infrastructures. The integration of map
production and field statistical survey can be
an important tool providing synergy to
improve the characteristics of the respective
products. The integration of the LUCAS
Eurostat project and Corine land cover offers
scope for validating future erosion modelling
efforts, considering that LUCAS could make
available field data on soil erosion
phenomena all over Europe.
The array of indicator characteristics would
offer a standardised approach that can be
repeated as data sets become more accurate

or more readily available. A method based on


both runoff and land cover would be
commendable, since both are measurable
and both significantly affect soil erosion.
USLE-based methods do not recognise the
physical rationale behind the erosion
process, and the relative importance of the
different factors is not always satisfactory.
However, the development of runoff based
indicators will require more time. Therefore
factor-based methods similar to the USLE
approach will have to be adopted in the
interim to produce a first workable solution.
Runoff and cover changes should be viewed
in conjunction with management practices.
Policy-makers may also prefer compatibility
with an international framework such as that
established within the OECD framework. It is
of no use to invent indicators that would
wrongly direct subsidies. Focus should be on
encouraging soil conservation strategies.

53

54

Assessment and reporting on soil erosion

7. General discussion on indicators


7.1. Data availability for soil erosion
indicators
An overview of existing databases available at
the JRC for the calculation of a soil erosion
indicator of state was given. The European
soil profile database (SPADE), containing
data on 355 soil profiles across western
Europe, should provide the necessary
information for assessing soil vulnerability to
erosion. Examples of pedotransfer rules for
soil erodibility and soil crusting show the
importance of incorporating baseline soil
data into soil erosion risk assessments.
However, it is suggested that erosion risk
assessment should be based on catchment
areas and not on soil polygons. The 500 to
1000 km European catchment geographical
database (scale 1:1 000 000) should suit the
purpose.

7.2. Indicators of state


An indicator of state to be developed in the
short term should be based on topography,
soil type and land cover. There is a perceived
shortcoming of high-resolution digital
elevation data on the European scale. Slope
classes derived from the European soil
database will therefore be the basic
topographic input. Pedotransfer rules for soil
crusting and erodibility were developed by
INRA-Orlans (Le Bissonnais and Daroussin,
2001) based on soil type, texture and parent
material. A second pedotransfer rule will
have to be derived for stoniness. The Corine
land use map should then be overlaid with
the slope classes, soil crustability and
stoniness maps in order to produce a first soil
erosion indicator of state. However, care
should be taken with the mixed nature of
cover and use in the Corine legend. There is
a need to have an agreement on the
operational procedure for state indicators
(and other indicators), for Member States to
follow.
The most challenging part will be the
development of the scoring system to the
different factors. Experts should regularly
upgrade the scoring system and explore its
various alternatives. Metadata should be
included on the confidence level of the
estimation.

In the longer term, improvements should


include climatic factors and regular updates
in land cover. Land cover updates should
make use of vegetation indices derived from
optical earth observation. The availability of
higher resolution DEMs across the EU is a
matter of urgency the current 1x1 km2 is
not adequate for erosion state indicators
The time aspects of more physically based
soil erosion indicators should be explored in
more detail.

7.3. Indicators of impact


A distinction ought to be made between offsite and on-site impacts.
Off-site indicators of impact deal with
sediment loads in rivers and freshwater
bodies. In the short term, agro-chemical
application at NUTS level 3 should be
overlaid with the erosion score developed in
the previous section. In the long term, data
on costs of sediment removal (from rivers,
canals, lakes and ditches) should be
collected. Another indicator to be developed
in the long term is water quality, whereby
distance and connectivity to freshwater
bodies will have to be considered in the offsite effects of soil erosion. However, care has
to be taken with the interpretation of off-site
indicators of impact (see Section 5.2).
Meaningful on-site indicators are more
difficult to develop. Crop productivity springs
to mind as a clear on-site effect. However,
relating actual yield to soil loss is extremely
difficult in European high-input agricultural
systems. Moreover, it will have to be
monitored over a long time period. In
Mediterranean regions, productivity should
be confronted with the available water
capacity calculated over soil depth. Loss of
soil fertility or soil quality could be
considered, but are difficult to measure.
Particularly for wind erosion, costs of reseeding could be estimated.
Data on land management changes (e.g.
tillage practices) and in particular
conservation practices should be collected in
a systematic manner. This could be realised
in conjunction with the IACS system. This
type of indicator accentuates the response
rather than the impact of soil erosion.

Part III Recommendations for further work

Part III Recommendations for


further work
8. Recommendations to the EEA
All recommendations are related to
accelerated soil erosion (i.e. where the
natural soil erosion rate has been
significantly increased by human activities
that cause changes in land cover and
management). In a first instance, the EEA
should focus on soil erosion by water and add
soil erosion by wind or tillage erosion in a
later phase.

8.2. Recommendations related to the


DPSIR assessment framework

The set of recommendations follow the main


chapters of the report. They relate to the
following categories: general, DPSIR
assessment framework, proposed indicators,
land use and soil erosion indicators, and
regional erosion assessment.

1. Although a policy-relevant integrated


assessment of soil erosion should not aim
at understanding or analysing soil
erosion as a process, the full range of
underlying factors that influence soil
erosion should be considered. These
factors include topography, soil, climate,
land cover (including vegetation), land
use and land management.

8.1. General recommendations


The following general recommendations are
related to the general reporting and
networking activities.
1. Since soil erosion has impacts on several
media, in particular on water quality,
working links should be developed with
other ETCs and specifically with the ETC
on Water. The Water Framework
Directive recognises the relevance of
agriculture as a major source of water
pollution.
2. Working links with groups of experts
contributing to the development of
international initiatives such as the
COST Action 623 on Soil erosion and
global change, the European Society for
Soil Conservation, IGBP-GCTE Focus 3
Soil erosion network (COST623 2001),
the Eurostat projects IRENA and LUCAS
should be maintained.
3. Institutional links with data providers
should be strengthened if the EEA is to
provide policy-makers and the general
public with information on the state of
the environment. A general complaint
was that data, and particularly statistical
data, exist but are often not accessible.

The DPSIR assessment framework is an


excellent approach onto which further
extensions and strategies of reporting on soil
erosion can be built. The following
recommendations are made to the EEA in an
attempt to extend the framework.

2. Particularly physical indicators should be


fully explored in the application of the
DPSIR assessment framework to soil
erosion and explicitly mentioned in the
resulting DPSIR scheme. Climate change
is considered as a driving force but only
in the sense that it relates to human
activities. Important physical factors that
influence soil erosion are topography,
soil type, soil vulnerability and climatic
factors (particularly rainfall). These
factors should not be separated from the
identified pressure indicators. At the
same time, headline indicators and subindicators should be identified and
prioritised.
3. All factors that change land cover, land
use and land management should be
included as driving forces. At present,
only agricultural intensification is seen as
the most important driving force (EEAETC/S, 1999; EEA, 2000). A revised
DPSIR scheme, presented in Figure 2.5,
has therefore been proposed, but could
certainly be elaborated further upon.
Examples of driving forces to be included
are human population, land
development, tourism, transport, natural
events and climate change.

55

56

Assessment and reporting on soil erosion

4. The revised DPSIR scheme (Figure 2.5)


presents land cover change and
precipitation as the most important
pressure indicators of soil erosion, as they
are seen to be directly influencing the
degree of soil erosion.
5. The DPSIR assessment framework lends
itself to systems analysis and as such is
very useful in describing the relationships
between the origins and consequences of
environmental problems. Obviously, the
real world is more complex than can be
expressed in simple causal relationships.
Linkages between the different types of
indicators are explored through the
DPSIR chain. However, the linkages
deserve further attention, not least to
capture the dynamics of the system.
Moreover, linkages within one type of
indicators (e.g. pressures) are not
explored, despite their repeatedly
reported importance.
6. There is a huge difference between
measured erosion, actual erosion risk
and potential erosion risk. Indicators
describing the driving forces and
pressures may affect the risk of soil
erosion, but they may not affect soil
erosion in itself at present, which also
depends on underlying physical factors
such as soil vulnerability and climatic
conditions. A mechanism is therefore
needed to jointly estimate the potential
and actual risk, based on links between
the identified driving force and pressure
indicators, and on an estimation or
measurement of what is actually
happening.
7. In the different reports made by the EEA,
it is recognised that a distinction ought to
be made between on-site and off-site
impacts of soil erosion. This distinction,
however, already applies at an earlier
stage in the DPSIR chain, namely at the
stage of state indicators. Soil erosion can
be measured in terms of actual sediment
loss per unit area (on-site) or in terms of
sediment delivery into streams or rivers
(off-site).
8. At present, there is no reporting
mechanism in place to assess whether
existing measures are leading to
improvement of soil conditions or to
gauge the level of implementation of
existing legislation. This could be a focal
point of action.

9. The current level of detail chosen for the


application of the DPSIR assessment
framework to soil erosion implicitly
enables the identification of broad
groups of actors related to the perceived
environmental problem. However, the
full identification of the several actors
involved requires a more detailed
stakeholder analysis, which ultimately
would help formulate sound policies for
remediation and mitigation strategies.

8.3. General recommendations


related to the proposed
indicators
Recommendations related to the indicators
proposed by EEA-ETC/S (1999, 2000) (see
also Table 2.1) are presented below. A
separate section is devoted to the indicators
of state (Section 8.5). A number of
recommendations are also provided that are
related to land use issues in the indicators for
soil erosion (see Section 8.4).
1. Driving forces, other than agricultural
intensification, should be included (see
above).
2. Driving forces or pressures should never
be evaluated alone in relation to erosion.
In order to understand the complexity of
accelerated erosion, it is necessary that at
least some of the indicators identify the
causes of soil erosion. Physical factors
that influence erosion rates are
topography, soils, climate and land cover.
Land cover is in turn influenced by the
socioeconomic environment and as such
by anthropogenic activities, notably land
use and management. Physical factors
should be explicitly mentioned and
linked to the existing indicators.
3. The six proposed pressure indicators
relate to agricultural intensification. It
should be made explicit that all are
complex and not directly linked to the
phenomenon of soil erosion. Moreover,
the indicators are usually only averages
on a large area basis and should
therefore be carefully interpreted.
4. GISCO databases, such as NUTS and
Corine land cover, can be used together
with farm structure survey (FSS) data
from Eurostat to derive the proposed
indicators of agricultural intensification.
However, a concise effort should be made
to spatialise or disaggregate the

Part III Recommendations for further work

agricultural statistical data to the


maximum possible.
5. The lack of good quality data on actual
fertiliser use makes it a weak indicator at
present.
6. In addition to crop yield, crop type, crop
rotations, crop management and area
devoted to a particular crop should be
considered as indicator of pressure.
7. The nature of soil erosion has to be
assessed in order to evaluate the on-site
loss and the possible off-site impacts. The
identified indicators of state and impact
are difficult or expensive to measure and
the data are usually not readily available.
An effort should be made to compile and
centralise existing data.
8. The area affected by erosion is the key
indicator for soil erosion, and should be
augmented with an indication of the
magnitude of erosion in a particular area.
A separate section is devoted to this set of
state indicators.
9. Indicators of state have to be a measure
of soil loss, and should explicitly relate to
climate, topography, soil properties, land
cover and land management.
10. The extent and severity of soil erosion
will have to be quantified and related to
land cover changes.
11. Sediment delivery ratio or sediment loads
in rivers are not all directly related to soil
erosion (part may be caused by riverbank
or channel erosion). There may be a time
lag between conservation measures and
sediment measurements so that care has
to be taken in the interpretation of
sediment concentrations as an indicator
of impact. Sedimentation (and the link to
eutrophication) of lakes is also important
and necessitates a link with water quality
studies.
12. Indicators of response are conservation
practices and mitigation strategies, which
are rarely in existence at present.
However, a concise effort should be made
to monitor prevention and control
measures.

8.4. Recommendations related to


land use and soil erosion
indicators
Land use and management are the result of
human activities and as such are the most
important factors that influence and control
accelerated soil erosion. Land cover may be
radically altered within a short time, but
physical and biological changes within the
soil, affecting erosion rates, may take longer
periods. The following recommendations
relate to the importance of considering land
use issues in the development of soil erosion
indicators.
1. As stated above, driving forces and
pressures should be expanded to all
factors that influence land cover.
2. A risk analysis is recommended in order
to highlight the risk that is specifically
related to the type of land cover. This
involves a distinction between actual and
potential soil erosion risk (see
recommendations for indicator of state).
3. Land cover type and change are the best
pressure indicators for soil erosion, as
they directly control the intensity and
frequency of overland flow and soil
erosion. Land cover type and changes,
including forest fires and deforestation,
can be detected by combining the
reference land cover database, Corine
land cover, with vegetation changes
indices from NOAA-AVHRR, SPOT
Vegetation or other earth observation
derived indices.
4. Precipitation regimes directly and
indirectly, through their influence on
land cover, influence soil erosion and
should therefore be included as
important pressure indicators. These
regimes can be detected using the
GISCO climate coverages and the
monitoring agriculture by remote
sensing (MARS) meteorological
database. The combination of
precipitation regimes with other physical
factors such as topography (e.g. aspect)
should also be considered.
5. Depending on the particular type of land
use and management, including
intensity, land resources are subject to a
given degree of stress. Land use and
management should therefore be
monitored as an important factor that

57

58

Assessment and reporting on soil erosion

influences soil erosion. Tillage erosion is


a prime example of human-induced
erosion.
6. Indicators should be developed for
monitoring the effectiveness and level of
enforcement of soil protection policies.

8.5. Recommendations related to


regional erosion assessment
(indicators of state)
The area affected by erosion is an important
indicator of state for soil erosion. Ultimately,
it is the area that is affected by soil erosion
and an estimate of the expected magnitude
in a particular area that policy-makers would
need to know in order to formulate a sound
soil protection policy. Regional soil erosion
assessment is therefore needed in order to
make objective comparisons that may provide
a basis for further environmental analysis,
economic statements or policy development.
Two important forms of erosion assessment
that reflect the current state of degradation
are measurements and field observations.
Apart from the time and expenses related to
collecting these types of distributed point
data, spatial interpolation is not justified due
to the sporadic distribution and episodic
occurrence of soil erosion. Regional soil
erosion assessment therefore requires other
techniques to be used taking care not to
neglect measurements and field
observations.
1. Effective monitoring and reporting on
soil erosion can only take place when the
following concepts related to soil erosion
are understood: (a) the fundamental
processes of soil erosion and in particular
of soil erosion by water as this is the most
important form of soil erosion, (b)
accelerated soil erosion and (c) the
underlying bio-physical and
socioeconomic factors that influence soil
erosion.
2. Indicators for soil erosion should be
developed according to the following
properties and procedures: quantitative,
objectively calculated, validated against
measurements and evaluated by experts.
3. Actual soil erosion measurements, such
as collected for the hot-spot map, should
continue to be compiled.

4. Field observations are invaluable as soil


erosion indicators of state. However, the
impossibility of making truly objective
comparisons between and often within
areas calls for a standardised approach to
record and particularly map the
observations.
5. In conjunction with soil erosion
measurements and observations, data on
climate, topography, soil and land use
should be carefully documented for each
observation or measurement. The
erosion type, scale of measurement/
observation, study period should be well
documented. This requires the set-up of
a comprehensive database, including
metadata.
6. A Europe-wide monitoring network for
soil such as proposed by the EEA (2001b)
should include soil erosion, covering the
most affected areas (hot spots). A
standardised approach to record soil
erosion should be defined.
7. Questionnaire-based mapping
approaches provide quick results for
creating awareness, but should be
avoided in the future whilst not rejecting
field observations and measurements.
8. The temporal and spatial patchiness of
soil erosion favours a risk analysis
approach in order to make comparisons
between regions and to complement
field measurements and observations.
The underlying model in a risk
assessment should ideally translate
experts knowledge into mathematical
algorithms. The widespread availability of
GIS data for key controlling variables
strongly favours a factorial or modelling
base for assessments of soil erosion.
9. Factorial models are useful for
identifying the extremes of low and high
erosion, but less satisfactory in
identifying the gradation between the
extremes. There are difficulties about
combining different factor ratings into a
single scale, about the individual
weightings and about the assumed
linearity and statistical independence of
the separate factors. A process modelling
approach is therefore recommended in
case the full spectrum of soil erosion has
to be assessed.

Part III Recommendations for further work

10. The difficulties associated with a process


modelling approach should not be
underestimated and a suitable model
should (a) represent the state of the art
in current understanding of soil erosion,
(b) respond explicitly and rationally to
changes in climate and land use, (c)
combine sufficient simplicity for
application on a regional scale and (d)
relate coarse scale forecasts to measured
erosion rate data so that explicit
validation can be made with field
monitoring data, to make full use of
experimental sites. The process
modelling method has the advantage of
producing an indicator of state with the
possibility for analysing different
scenarios, which in turn enables the
formulation of soil conservation policies.
The Pesera project has adopted this
modelling approach.
11. Modelling efforts should be thoroughly
validated against erosion measurements,
and a clear distinction should be made

between modelled erosion risk and


present-day erosion rates.
12. A programme to monitor soil erosion
across different agro-ecological regions
and under different land uses should
underpin both mapping exercises and
regional soil erosion risk assessment
methods. Only then a sound approach is
ensured of estimations and mapping
features that are directly validated and
compared with measurements. Moreover,
measuring campaigns may lead to new
insights and therefore to both better
mapping and risk assessments.
13. Erosion literature commonly identifies
tolerable rates of soil erosion, but these
rates usually exceed the rates that can be
balanced by weathering of new soil from
parent materials, and can only be
considered acceptable from an economic
viewpoint. Tolerable soil loss rates should
be developed but at the same time
carefully evaluated by experts.

59

60

Assessment and reporting on soil erosion

9. Conclusions
The DPSIR assessment framework is an
excellent tool onto which further extensions
and strategies of reporting can be built. A
revised scheme for erosion within the
framework presents changes in land cover
and precipitation as the most important
pressure indicators of soil erosion. The
DPSIR assessment framework sets a good
basis for identifying the different factors
influencing soil erosion, but, at the current
level of detail, the resulting scheme for soil
erosion does not explicitly allow for the full
identification of actors in the DPSIR chain.
Driving forces and related pressure
indicators other than agricultural
intensification should be included. However,
their relationship with soil erosion is
complex. Physical factors that cause erosion
should be included, i.e. topography, soils,
climate and land cover, and their interaction
with pressures should be analysed. The
identified indicators of state and impact are
difficult or expensive to measure and the
data are usually not readily available.
Indicators of response are prevention and
control measures, which are rarely in place at
present.
Generally, it was concluded that the
indicators should be developed according to
the following properties and procedures:
quantitative, objectively calculated, validated
against measurements and evaluated by
experts.
Land cover type and change, land
management and land use are the best
pressure indicators for soil erosion. Land
cover type and change can be monitored by
combining Corine land cover data with earth
observation derived indices. In addition, land
use and management information can be
derived from Eurostat, together with the
farm structure survey data. The statistical

data should be spatialised and disaggregated


to the maximum possible.
A regional assessment using a combination of
modelling, expert estimates and other
methods should be developed in order to
provide a general view and identify the hotspot areas where to undertake a detailed soil
erosion monitoring programme.
Regional soil erosion assessments enable
estimates of the area that is affected by soil
erosion and the expected magnitude in a
particular area, both of which are required to
formulate sound soil protection policies.
Indicators of state should reflect all four
strategies of regional soil erosion assessment,
i.e. distributed point data, expert mapping,
factor mapping and process modelling. The
four different methods described in this
report are not mutually exclusive and each
provides a different emphasis. Erosion rate
measurements and field observations provide
an unambiguous measure of actual erosion,
where they exist. However, apart from the
time and expenses involved, spatial
interpolation is not justified due to the
sporadic distribution and episodic
occurrence of soil erosion. Factorial
approaches provide a measure of erosion risk
and can only be recommended for
identifying the extremes of low and high
erosion, but not for the gradation between
the extremes.
A process modelling method is
recommended for modelling soil erosion risk
in relation to climate and land use changes.
Field campaigns are necessary and databases
should be made with erosion measurements,
field observations, data on underlying factors
influencing erosion (climate, topography,
soils and land use) and related metadata
(period of record, erosion type, etc.).

Part III Recommendations for further work

10.References
Alcamo, J. (ed.) (1994). IMAGE 2.0: Integrated
modelling of global climate change. Kluwer,
London.
Arnoldus, H. M. J. (1978). An
approximation of the rainfall factor in the
universal soil loss equation. In: De Boodt, M.
and Gabriels, D. (eds). Assessment of erosion,
Wiley, Chichester, pp. 127132.
Batjes, N.H., 1996. Global assessment of land
vulnerability to water erosion on a by
degree grid. Land Degradation & Development,
7:353-365.
Batjes, N. H. (2000). Degradation and
vulnerability database for central and eastern
Europe Preliminary results of the Soveur project.
Proceedings of concluding workshop
(Busteni, 2631 October 1999), FAO, Rome,
RISSA, Bucharest, and ISRIC, Wageningen,
pp. viii + 99.
Blum, W.E.H. 1998. Soil degradation caused by
industrialization and
urbanization. In: Blume H.-P., H. Eger, E.
Fleischhauer, A. Hebel, C. Reij, K.G. Steiner
(Eds.): Towards Sustainable Land Use, Vol. I,
755-766, Advances in Geoecology 31,
Catena Verlag, Reiskirchen.
Boardman, J. (1998). An average soil erosion
rate for Europe: Myth or reality?, Journal of
Soil and Water Conservation 53(1), 4650.
Boardman, J. and Favis-Mortlock, D. T. (eds)
(1998). Modelling soil erosion by water, SpringerVerlag NATO-ASI Series I-55, Berlin. 531 pp.
Brazier, R. E., Rowan, J. S., Anthony, S. G. and
Quinn, P. F. (2001). Mirsed: Towards an MIR
approach to modelling hillslope soil erosion
at the national scale, Catena 42, (1) 5979.
Briggs, D. J. and Giordano, A. (1995). Corine
soil erosion report, European Commission, 124
pp.
BTG, 1998. Bridging the gap conference, 3 to 5
June, 1998 Chairmans conclusions New needs

and perspectives for environmental information,


EEA/UK/NL, 1998.
Burt, T. P. (1994). Long-term study of the
natural environment Perceptive science or
mindless monitoring?, Progress in Physical
Geography 18(4), 475496.
CEC (1985). Explanatory text and map sheets of
the 1: 1.000.000 soil map of the European
Communities. Directorate-General for
Agriculture. Office for official publications of
the European Communities, Luxembourg,
Luxembourg.
Corine (1992). Corine soil erosion risk and
important land resources in the southern regions of
the European Community, Publication EUR
13233 EN, Luxembourg, 97 pp. + maps.
COST623 (Soil erosion and global change)
(2001). Mid-term report on research cooperation in
Europe, report to European Commission.
De Jong, S. M. (1994). Applications of
reflective remote sensing for land
degradation studies in a Mediterranean
environment, Nederlandse Geografische Studies
177.
De Jong, S. M., Brouwer, L. C. and Riezebos,
H. T. (1998). Erosion hazard assessment in
the Peyne catchment, France. Working
paper DeMon-2 project. Dept. Physical
Geography, Utrecht University.
De Ploey, J. (1989). A soil erosion map for
western Europe. Catena Verlag.
De Roo, A. P. J. (1993). Modelling surface
runoff and soil erosion in catchments using
geographical information systems; validity
and applicability of the Answers model in
two catchments in the Loess area of south
Limburg (the Netherlands) and one in
Devon (UK), Nederlandse Geografische Studies
157.
Desmet, P.J.J. and G. Govers, 1996. A GIS
procedure for automatically calculating the
USLE LS factor on topographically complex
landscape units. Journal of Soil and Water
Conservation 51: 427-433.

61

62

Assessment and reporting on soil erosion

EEA (1995). Europes environment: The Dobris


assessment, European Environment Agency,
Copenhagen, Denmark. 676 pp.
EEA (1998). Europes environment: The second
assessment. Environmental Assessment Report
No 2. European Environment Agency. 295
pp.
EEA (1999a). Environment in the European
Union at the turn of the century. European
Environment Agency.
EEA (1999b). Environmental indicators:
Typology and overview. European Environment
Agency.
EEA (2000). Down to earth: Soil
degradation and sustainable development in
Europe, Environmental Issues Series, Number
16, 32 pp. European Environment Agency.
EEA (2001a). Analysis and mapping of soil
problem areas (hot spots) in Europe. Final
Report to EEA, prepared by Turner, S.,
Lyons, H. and Favis-Mortlock, D. In: Where are
the hot spots of soil degradation in Europe?, CDROM distributed to EIONET for review.
European Environment Agency.
EEA (2001b). Proposal for a European soil
monitoring and assessment framework, Technical
Report No 61/2000, prepared for the
EIONET soil workshop in Vienna, 1214
October 1999, Vienna. European
Environment Agency.
EEA (2001c). European soil monitoring and
assessment framework. EIONET workshop
proceedings, Technical Report No 67/2001,
European Environment Agency.
EEA (2002a). EIONET workshop on indicators
for soil contamination. Workshop proceedings.
Technical Report No 78/2002, European
Environment Agency.

EEA (2002b). Proceedings of the technical


workshop on indicators for soil sealing.
Technical Report No 80/2002, European
Environment Agency.
EEA-ETC/S (1999). Final report on Task 6 of
the Technical Annex for the subvention to the
European Topic Centre on Soil. Working report
prepared by Dwel, O. and Utermann, J.
European Commission (2001). The sixth
environmental action programme, COM(2001)
31 final. 2001/0029 (COD).

European Commission (2002). Towards a


strategy for soil protection, COM(2002) 179
final. Internet: http://europa.eu.int/comm/
environment/agriculture/
soil_protection.htm.
'Favis-Mortlock, D., and J. Boardman (1999)
Soil Erosion Hot Spots in Europe. Unpublished
report for the EEA, Copenhagen.
Fournier, (1972). Soil conservation, Nature
and Environment, No 5, Council of Europe,
Strasbourg.
GCTE (1997). GCTE Focus 3 erosion network:
1997 model, experimental and monitoring
metadata, Global Change and Terrestrial
Ecosystems Report No 6, GCTE Focus 3
Office, Wallingford, UK. 136 pp.
Gentile, A. R. (1999a). From national
monitoring to European reporting The EEA
strategy and activities on soil related issues,
presented at the preparatory meeting of the
first European Soil Forum.
Gentile, A. R. (1999b). Towards the development
of a system of policy-relevant indicators on soil.
European Soil Forum, Berlin 1999.
Gobin, A. and Govers, G. (2001). First annual
report of the pan-European soil erosion risk
assessment project.
Gobin, A. and Govers, G. (2001). Pesera (panEuropean soil erosion assessment) EC Contract No
QLK5-CT-1999-01323, first interim report,
69pp.
Gobin, A. and Govers, G. (2002). Second
annual report of the pan-European soil erosion risk
assessment project.
Gobin, A., Govers, G., Kirkby, M., Jones, R.,
Kosmas, C., Puigdefabregas, J., Van Lynden,
G. and Le Bissonnais, Y. (1999). Technical
Annex: Pan-European soil erosion risk assessment
project.
Greenland, D. J. and Szabolcs, I. (1994).
Proceedings of the Soil Resilience and Sustainable
Land Use Symposium. Greenland, D. J. and
Szabolcs, I. (eds), CAB International, 1994.
Harrod, T. R. (1994). Runoff, soil erosion
and pesticide pollution in Cornwall. In:
Rickson, R. J. (ed.), Conserving soil resources:
European perspectives, CAB International,
Wallingford, UK. pp. 105115.

Part III Recommendations for further work

Hasholt, B. (1988). On the assessment of soil


erosion in Denmark. In: Morgan, R. P. C.
and Rickson, R. J. (eds), Agriculture: Erosion
assessment and modelling, Commission of the
European Communities, Publication EUR
10860, Brussels, Belgium. pp. 5572.
Hasholt, B. (1998). Assessment of erosion
and some implications for model validation.
In: Summer, W., Klaghofer, E. and Zhang, W.
(eds), Modelling soil erosion, sediment transport
and closely related hydrological processes, IAHS
Press Publication No 249, Wallingford, UK.
pp. 249-260.

Mairota, P., Thornes, J. B. and Geeson, N.


(1997). Atlas of Mediterranean environments in
Europe: The desertification context, Wiley,
Chichester, UK, 224 pp.
Mircea, M. (1983). The average rate of soil
degradation by erosion in Romania,
Informative Bulletin of the Academy of
Agricultural and Forestry Sciences, Bucharest,
No 12, pp. 4765.
Mitchell, J. K. and Bubenzer, G. D. (1980).
Soil loss estimation. In Kirkby and Morgan
(eds). Soil erosion, Wiley and Sons, Chichester,
New York.

Ingram, J.S.I., Lee, J.J. and Valentin, C.


(1996). The GCTE Soil Erosion Network: a
multi-participatory research program. Journal
of Soil and Water Conservation 51(5), 377-380.

Montier, C., Daroussin, J., King, D. and Le


Bissonnais, Y. (1998). Cartographie vde lala
Erosion des sols en France. INRA, Orlans.

Jger, S. (1994). Modelling regional soil


erosion susceptibility using the universal soil
loss equation and GIS. In: Rickson, R. J
(ed.). Conserving soil resources: European
perspectives, CAB International, pp. 161177.

Morgan, R. P. C, Morgan, D. D. V. and Finney,


H. J. (1984). A predictive model for the
assessment of soil erosion risk, Journal of
Agricultural Engineering Research 30, p. 245
253.

King, D., Stengel, P. and Jamagne, M. (1999).


Soil mapping and soil monitoring: State of
progress and use in France. In: Bullock, P.,
Jones, R. J. A. and Montanarella, L. (eds). Soil
resources of Europe. EUR 18991 EN, 204 pp.,
Office for Official Publications of the
European Communities, Luxembourg.

Morgan, R. P. C. (1992). Soil erosion in the


northern countries of the European Community.
EIW Workshop: Elaboration of a framework of a
code of good agricultural practices, Brussels, 21
22 May 1992.

Kirkby, M. J. and King, D (1998). Summary


report on provisional RDI erosion risk map
for France. Report on contract to the
European Soil Bureau (unpublished).
Kirkby, M. J., Le Bissonais, Y., Coulthard, T. J.,
Daroussin, J. and McMahon, M. D. (2000).
The development of land quality indicators
for soil degradation by water erosion,
Agriculture, Ecosystems and Environment 81,
12536.
Kosmas, C., Kirkby, M. J., and Geeson, N.
(eds) (1999). The Medalus project: Manual on
key indicators of desertification and mapping of
environmentally sensitive areas to desertification.
European Commission, EUR 18882, 87 pp.
Le Bissonnais, Y. and Daroussin, J. (2001). A
pedotransfer rules database to interpret the soil
geographical database of Europe for environmental
purposes. Proceedings of the workshop on the
use of pedotransfer in soil hydrology
research in Europe, Orlans, France, 1012
October 1996.

Morgan, R. P. C. (1995): Soil erosion and


conservation. Second Edition. Longman,
Essex.
Nearing, M. A., Foster, G. R., Lane, L. J. and
Finkner, S. C. (1989). A process-based soil
erosion model for USDA-water erosion
prediction project technology, ASAE Trans
32, 15871593.
OECD (1993): OECD core set of indicators
for environmental performance reviews.
Environment Monographs, No 83, Paris.
OECD (1999). Environmental indicators for
agriculture: Concepts and frameworks, Vol. 1.
Organisation for Economic Cooperation and
Development, Paris.
Oldeman, L. R., Hakkeling, R. T. A. and
Sombroek, W. G. (1991). Glasod world map of
the status of human-induced soil degradation
(second revised edition), ISRIC,
Wageningen, UNEP, Nairobi.
Pimentel, D., Harvey, C., Resosudarmo, P.,
Sinclair, K., Kurz, D., McNair, M., Crist, S.,
Shpritz, L., Fitton, L., Saffouri, R. and Blair,

63

64

Assessment and reporting on soil erosion

R. (1995). Environmental and economic


costs of soil erosion and conservation
benefits, Science 267, 11171123.
Poesen, J., Vandaele, K. and Van Wesemael,
B. (1996a). Contribution of gully erosion to
sediment production on cultivated lands and
rangelands. In: Walling, D. E. and Webb, B.
W. (eds), Erosion and sediment yield: Global and
regional perspectives, IAHS Publication No 236,
Wallingford, UK. pp. 251266.
Poesen, J., Boardman, J., Wilcox, B. P. and
Valentin, C. (1996b). Water erosion
monitoring and experimentation for global
change studies, Journal of Soil and Water
Conservation 51(5), 386390.
Poesen, J., Vandaele, K. and Van Wesemael,
B. (1998). Gully erosion: Importance and
model implications. In: Boardman, J. and
Favis-Mortlock, D., Modelling soil erosion by
water, Springer-Verlag Berlin Heidelberg.
Renard, K. G., Foster, G. R., Weessies, G. A.,
McCool, D. K. and Yoder, D. C. (eds) (1997).
Predicting soil erosion by water: A guide to
conservation planning with the revised universal
soil loss equation (RUSLE). US Department of
Agriculture, Agriculture Handbook 703.
RIVM (1992). The environment in Europe: A
global perspective. Report 481505001, RIVM,
Bilthoven, the Netherlands.
Stanners, D. and Bourdeau, P. (eds) (1995).
Europe's Environment, European Environment
Agency, Copenhagen, Denmark
UNCED, 1993. UNCED, 1993. Agenda 21:
Programme of action for sustainable development.
Proceedings UN Conference on
Environment and Development, New York.
Van der Knijff, J. M., Jones, R. J. A. and
Montanarella, L. (1999). Soil erosion risk
assessment in Italy, EUR 19044 EN, 52 pp.
Van der Knijff, J. M., Jones, R. J. A. and
Montanarella, L. (2000). Soil erosion risk
assessment in Europe, EUR 19044 EN, 34 pp.

Van Engelen, V.W.P. and Wen, T.T., (eds.),


1995. Global and national soils and terrain
databases (SOTER). Procedures manual
(revised edition). UNEP-ISSS-ISRIC-FAO,
Wageningen. 125 p.
Van Lynden, G. W. J. (1995). European soil
resources, Nature and Environment, No 71.
Council of Europe, Strasbourg.
Van Lynden, G. W. J. and Oldeman, L. R.
(1997). The assessment of the status of humaninduced soil degradation in south and south-east
Asia (ASSOD). ISRIC, Wageningen, the
Netherlands.
Van Muysen, W., Govers, G., Bergkamp, G.,
Poesen J. and Roxo, M. (1999).
Measurement and modelling of the effects
of initial soil conditions and slope gradient
on soil translocation by tillage, Soil and
Tillage Research 51: 303.
Wischmeier (1975): in Stewart, B. A. et al.,
Control of water pollution from cropland. Vol. I, A
manual for guideline development. Agricultural
Research Service, Hyattsville, MD.
Wischmeier (1976): in Stewart, B. A. et al.,
Control of water pollution from Cropland. Vol. II,
An overview. Agricultural Research Service,
Washington DC.
Wischmeier, W. H. and Smith, D. D. (1978).
Predicting rainfall erosion losses A guide
for conservation planning, Agriculture
Handbook 537, US Department of
Agriculture.
Woolhiser, D. A., Smith, R. E. and Goodrich,
D.C. (1990). Kineros, A kinematic runoff
and erosion model: Documentation and user
manual. USDA Agricultural Research
Service, ARS-77, pp. 130.
Yassoglou, N., Montanarella, L., Govers, G.,
Van Lynden, G., Jones, R. J. A., Zdruli, P.,
Kirkby, M., Giordano, A., Le Bissonnais, Y.,
Daroussin, J. and King, D. (1998). Soil erosion
in Europe. European Soil Bureau.

Annexes

Annexes
Annex I List of participants
Participants

Address

Anna Rita Gentile


Anna.Rita.Gentile@EEA.eu.int

European Environment Agency


Kongens Nytorv 6, DK-1050 Copenhagen K
Tel. (+45) 33 36 71 00; Fax (45) 33 36 71 99

Anne Gobin
Anne.gobin@geo.kuleuven.ac.be

Laboratory for Experimental Geomorphology


Redingenstraat 16, B-3000 Leuven
Tel. (32-16) 32 64 33; Fax (32-16) 32 64 00

Gerard Govers
Gerard.govers@geo.kuleuven.ac.be

Laboratory for Experimental Geomorphology


Redingenstraat 16, B-3000 Leuven
Tel. (32-16) 32 64 33; Fax (32-16) 32 64 00

Paul Campling
Paul.campling@agr.kuleuven.ac.be

Ground for GIS


Vital Decosterstraat 102, B-3000 Leuven
Tel. (32-16) 32 97 32; Fax (32-16) 32 97 00

Simon Turner
Simon.Turner@adas.co.uk

ADAS, Wergs Road, Woodthorne, Wolverhampton WV6 8TQ, United


Kingdom

Hester Lyons
Hester.Lyons@adas.co.uk

ADAS, Wergs Road, Woodthorne, Wolverhampton WV6 8TQ, United


Kingdom

Luis Carazo
Luis.Carazo_jimenez@cec.eu.int

Commission of the European Communities


Unit B1, Water, the Marine and Soil
BU-9 Office No 3/137, Rue de la Loi/Wetstraat 200 B-1049 Brussels
Tel. (32-2) 2960 066

Pierpaolo Napolitano
napolita@istat.it

ISTAT DISS
Via A. Rava 150, I-00142 Rome
Tel. (39-06) 59 52 43 47; Fax (39-06) 59 43 257

Philipp. Schmidt-Thom
philipp.schmidt-thome@gsf.fi

PO Box 96
FIN-02151 Espoo
Tel. (Mobile) (358-40) 54 24 192
Tel. (Office) (358-20) 55 02 163
Fax (358-20) 55 012

Robert Jones
Robert.jones@jrc.it

Commission of the European Communities DG Joint Research


Centre
Environment Institute, European Soil Bureau
TP 262, Via E. Fermi, I-21020 Ispra (VA)
Tel. (39-332) 78 63 30; Fax (39-332) 78 99 36

Godert Van Lynden


Vanlynden@isric.nl

International Soil Reference and Information Centre


PB 353, Duivendaal 7-9, 6700AJ Wageningen
The Netherlands
Tel. (31-317) 47 17 15; Fax (31-317) 47 17 00

Olaf Dwel
Olaf.Duewel@bgr.de

Bundesanstalt fur Geowissenschaften und Rohstoffe Federal Institute


for Geosciences and Natural Resources
Stilleweg 2, D-30655 Hanover
Tel. (49-511) 64 32 841; Fax (49-511) 64 33 662

David Favis-Mortlock
d.favis-mortlock@qub.ac.uk

Queens University of Belfast, School of Geography Belfast BT7 1NN,


Northern Ireland
Tel. (44-28) 90 32 51 33; Fax (44-28) 90 32 12 80

Robert Evans
R.Evans@anglia.ac.uk

Department of Geography
Anglia Polytechnic University
East Road, Cambridge CB1 1PT
United Kingdom
Tel. (44-1223) 36 32 71 x 2662

65

66

Assessment and reporting on soil erosion

Annex II Agenda
Tuesday 27/3/2001
Convenor: Gerard Govers
Time

Speaker

Subject

Soil erosion indicator framework


14.0014.30

Anna Rita Gentile

General introduction and European framework applied to soil


erosion

14.3015.00

Olaf Dwel

Soil erosion indicator work at the former ETC/Soil

15.0015.30

Paul Campling and Costas


Kosmas

GISCO databases and tools to derive pressure indicators for


soil erosion

15.3017.00

Chair: Gerard Govers

Discussion on indicator framework

Wednesday 28/3/2001
Convenor: Gerard Govers
Time

Speaker

Subject

Regional / Spatial indicators


10.0010.30

Godert Van Lynden

The Glasod map

10.3011.00

Dave Favis-Mortlock

The hot-spot map

11.0011.30

Anne Gobin and Mike


Kirkby

Regional assessment of the impact of soil erosion by water

11.3013.00

Chair: Gerard Govers

Discussion on regional/spatial soil erosion indicators

13.0014.00

Lunch

Data availability
14.0014.30

Robert Jones

Data availability for soil erosion indicators at the European level

14.3015.30

Chair: Gerard Govers

Discussion on data availability, data gaps and needs

General discussion
15.3016.30

Chair: Gerard Govers

General conclusions and recommendations

16.3017.00

Anna Rita Gentile

Concluding remarks

Annex II Agenda

1. What is soil erosion?

A specific point of discussion was the


indicator of state for soil erosion. Soil erosion
is recognised to be highly variable in both
space and time. The following questions were
used as guidelines to discuss the assessment
of soil erosion.

2. What information does a policy-maker


need to assess soil erosion and its current
impacts, and to formulate remedial
measures in Europe?

1. Which erosion types should be


considered? (wind, water, gully, mass
movements, active versus non-active
erosion (old gullies)

3. Is the conceptual framework (DPSIR;


MF-MI approach) adequate to describe
soil erosion in Europe (its state, impacts
on the soil resource and on other media,
the causes and measures)?

2. How can or should tillage erosion be


incorporated in the framework?

4. Is the list of proposed indicators for soil


erosion adequate? How many and which
type of indicators should be advocated?
(ideas for change). For each indicator in
the list:
Is the indicator adequate?
Are the data used adequate?
Are the conclusions and is the
assessment correct?
What else should be taken into
account?

4. What should be the preferred scale for


assessing soil erosion taking into account
its use for policy-makers? (nested
strategies at multiple scales, etc.)

Questions to guide the discussions


In the evaluation and discussions during the
workshop, the following questions were used
to guide the review.

5. What are the driving forces of soil


erosion? (with specific attention to
agriculture)
6. Are the drivers of soil erosion sufficiently
known and how do they link to the
phenomenon?
7. Are there other quantifiable indicators of
impact apart from the proposed
indicator removal of sediment deposits?
8. Is the assessment of soil erosion in
Europe correct? Are the methods used
scientifically sound?
9. What are the recommendations for
further work?

3. Can thresholds be derived for policy


purposes? How should these be set?

5. How should the extent of the erosion


problem be mapped in relation to the
severity or frequency?
6. What are good indicators of state and
impact?

67

68

Assessment and reporting on soil erosion

Annex III Background papers


presented at the workshop
State of play of EEA work on soil
erosion indicators

soil erosion and EEA expectations from the


workshop were also discussed.

Anna Rita Gentile


Project Manager for soil and contaminated sites
European Environment Agency

In particular, the EEA organised the


workshop with the aim to take stock of the
work done, get expert advice on how to
proceed with the work on soil erosion,
connect with other relevant initiatives on soil
erosion at the European and national level
and help to define the work plan of the new
European Topic Centre on Terrestrial
Environment.

This first presentation provided some


background information and focused on the
objectives of the EEA work programme on
soil and a description of the European soil
monitoring and assessment framework. The
state-of-play of EEA work on indicators for

A selection of overheads is included below.

So far, the EEA has collected information


based on the best available data. However,
this approach, although allowing for the
provision of timely information, has shown
some limitations. For example, it may not
help rationalise ongoing data collection and
monitoring activities at the national and
European levels, possibly covering subjects
that are not needed, while resources should
be better employed to fill data gaps in other

priority areas (BTG, 1998). In order to help


streamline monitoring, assessment and
reporting activities, a broader approach is
required. In the long term, the objective is to
focus on the best needed data. This shift
should be obtained by building stronger links
to EU policy needs, by focusing on the
assessment of the environmental impacts of
soil degradation and by undertaking a more
detailed analysis in hot-spot areas.

Annex III Background papers presented at the workshop

69

70

Assessment and reporting on soil erosion

The presentation continued with the


illustration of the results achieved in the
development of the work on soil erosion (see
also Dwel and Utermanns presentation

later in this annex). Finally, the EEA needs in


terms of data and expertise were discussed.
These issues are described by a selection of
overheads below.

Annex III Background papers presented at the workshop

Soil erosion hot-spot map for Europe


David Favis-Mortlock
School of Geography
Queens University Belfast
Background
The map of soil erosion hot spots for Europe
(Figure 4.3 in main text: EEA, 2000; EEA,
2001) aims to identify problem areas for soil
erosion by water and wind. It is based upon
published observations and measurements of
erosion in the field, and is a development of
a map which was previously produced for the
EEA (Favis-Mortlock and Boardman, 1999).
An empirical approach was adopted in
preference to one based upon estimates from
erosion models such as that of Corine
(1992) since current models have
difficulty in dealing with erosions spatial and
temporal variability (Boardman and FavisMortlock, 1998. This is particularly true for
estimates of erosion over large areas, such as
Europe.
However, problems of over-generalisation
(Boardman, 1998) can also beset empirically
based studies of erosion rates over large areas
(e.g. Pimentel et al., 1995). For this reason,
an attempt has been made here to indicate
erosion at a hierarchy of scales (Table 1). The
two coarser levels of the hierarchy are purely

71

qualitative, whereas the finest (location)


level is associated with locally measured data
for erosion rates. Its empirical basis means
that areas and rates on this map refer to
actual erosion (i.e. erosion under present
land use and climate) rather than potential
(i.e. hypothetical) erosion. This is a clear
advantage for the formulation of soil
conservation policy.
However, relatively little reliable measured
data on soil erosion exist. This is the
principal disadvantage of the approach used
for this map. It is probably for this reason
that a majority of published maps of erosion
at the national scale are partially or wholly
based upon model results. This scarcity of
measured data appears to be the case for
both western and eastern Europe. For
example, a map of erosion in Romania
(Mircea, 1983) makes use of both model
results and measurements (Ion Ionita,
personal communication, 2000); however, to
disentangle these is far from simple.
The hierarchical approach used for this map
also renders it different from other
empirically based large-area erosion maps
such as Glasod (Oldeman et al., 1991) and
Map 7.3 (Water erosion of soils in Europe)
in Europes environment: The Dobr assessment
(EEA, 1995). Both these maps identify areas
Spatial units used to construct the hot-spot map

Unit

Description

Source of information

Zones

Broad zones where the nature of erosion is, in


general, similar

Expert opinion of the map compilers

Hot-spot areas

Hot-spot areas, mostly within these zones

Opinion of several experts via the map


of De Ploey (1989) plus that of the map
compilers

Locations

Locations, mostly within these hot-spot areas, for


which there are measurements of erosion rates

Published literature (see References)

Table A3.1

72

Assessment and reporting on soil erosion

with a subjectively similar severity of erosion,


irrespective of the conditions which
produced this erosion, and irrespective of its
wider impacts. By contrast, the approach
used here allows:
hot-spot areas to be grouped or
differentiated according to which zone(s)
they are in, i.e. on the differences or
similarities in erosional conditions;
quantification of erosion rates for points
within the hot-spot area (where data
permit).
Thus the reader can pick out broad
similarities and differences in terms of
both causes and impacts for those regions
of Europe which suffer from an erosion
problem; and can link these broad zones with
measured rates. With knowledge of the ways
in which affected regions resemble and differ
from each other, future European soil
conservation strategies can be more
efficiently targeted. For example, selection of
appropriate sites for remediation might
involve criteria such as:
is currently active (i.e. does not result
mainly from erosion in the past);
gives rise to off-site effects which are
significant in the short term, such as
flooding and water quality problems;
is likely (in the longer term) to experience
a significant on-site drop in agricultural
productivity as a result of erosion.
Data quality issues
The spatial and temporal occurrence of erosion
Erosion is patchy in both space and time (cf.
Figure 1). Loss of soil can be highly variable
even in areas of severe erosion. For water
erosion for example, the vagaries of
topography concentrate erosive flows so that
severe erosion in one field can be found side
by side with almost untouched areas.
Similarly, several years can pass between
major erosion events (water or wind) even in
erosion-prone regions. Long time series of
measurements of erosion are therefore
required to adequately estimate erosion
rates. Temporal distributions of erosion are
highly skewed, so that calculation of longterm average values for erosion is statistically
dubious (use of the median is preferable, but
uncommon).

(8)
(9)

Precise delineation of erosion hot spots is


therefore futile. Additionally, even within an
area which is designated as a hot spot it may
well be that only a minority of fields will show
obvious erosion at any time. Also since almost
all erosion monitoring studies operate for
only a relatively short period, any assessment
of erosion rates for these hot-spot areas is
fraught with uncertainty.
Spatial considerations regarding data collection
To a large extent (8), erosion is independent
of national boundaries. However, field
measurements of soil erosion may be
obtained during a study which is funded or
sponsored by a particular country, or by a
scientist who works within well-defined
regional boundaries. The emphasis placed
upon erosion studies also varies markedly
from country to country. As a result, the
availability of data on European erosion
varies strongly from region to region. There
is thus some risk both of spurious hot spots
being generated simply by the presence of
abundant data for an area, and also of the
inverse problem: lack of data resulting in
under-emphasis of an areas erosion
problems. Any Europe-wide study of erosion
must therefore exercise discrimination in the
face of possibly artefactual positive or
negative hot spots.
Techniques of data collection
Techniques of data collection are an issue
with respect to the erosion rates quoted here.
Even for the same location, erosion rates
obtained by different methods (9) are likely
to vary. This study has unavoidably had
to draw upon data for erosion rates which
were obtained by a range of methods. While
in some cases it is possible to reconcile such
methodological variations, in general the
result is to increase the uncertainty associated
with rates assigned to mapped hot spots. Soil
loss rates calculated from plot-sized areas
(the most common among the studies
reported here) can be up to one or two
orders of magnitude higher than sediment
yields calculated from catchments. However,
results from small areas such as plots do not
include the contribution which talweg
(valley-bottom) gullying can make to total
erosion: this may be over 40 % in north
Europe and over 80 % in south Europe
(Poesen et al., 1996a). Due to erosions
temporal variability, soil loss rates from single
events are generally not reported here,

Except where trans-border land use is strongly influenced by differing national policies.
For example, by field survey of rill depths, collection of sediment lost from a plot, or aerial photography.

Annex III Background papers presented at the workshop

except where measured data are scarce, e.g.


eastern Europe.
Other issues regarding data
Most of the source publications for this map
are in English. This is a definite limitation,
although ameliorated to some extent by the
use of English publications which summarise
earlier non-English work.
The design and use of this map
As described previously, this map shows
erosion on a three-level spatial hierarchy. For
proper use of this map it is vital to remember
the following caveats:
hot-spot boundaries are rather arbitrary;
even within a hot-spot area, erosion occurs
patchily;
there is a considerable variability and
uncertainty associated with all cited rates of
erosion;
expert judgment has played a major role in
the methodology used here. This is
unavoidable, given the complex nature of
erosions occurrence and the limitations of
currently available data. Thus it is
important to note that, just because erosion
is not indicated at a particular location on
this map, this does not imply that no erosion
occurs there. For example, erosion occurs
regularly in Denmark (Hasholt, 1988,
1998) but does not appear to be a major
problem there.
Boundaries for water and wind erosion hotspot areas in western and southern Europe
are in most cases modified from De Ploey
(1989), while others have been deduced
from the publications cited. Those for
eastern Europe are also derived from
individual publications, interpolated as
necessary.

73

Loess zone, and an eastern zone (10). In the


southern zone, severe water erosion results
from intense seasonal rainfall. This is often
associated with overgrazing or a move away
from traditional crops. Erosion here may be
of considerable age. The principal impact is
on-site: soil productivity decreases as a result
of thinning. The northern zone has
moderate rates of water erosion. This mostly
results from less intense rainfalls falling on
saturated, easily erodible soils. There is also
local wind erosion of light soils. Impacts here
are mainly off-site, as agricultural chemicals
from the norths more intensive farming
systems are moved into water bodies along
with eroded sediment. Partially overlapping
these two zones is the eastern zone, where
former large state-controlled farms produced
considerable erosion problems. Eroded
sediments here may also be contaminated
from former industrial operations. Other,
relatively minor, areas of erosion occur
outside these zones. Within all three zones,
there are hot-spot areas where erosion is
more serious. The coverage of reliable
measurements of erosion is very patchy, and
to an extent reflects the activities of
particular workers rather than the severity of
the problem.
Rates of erosion
As noted in the methodology, regional rates
inferred from this map must be very
tentative. None the less there is some
indication that average rates of soil loss are
higher in southern and eastern Europe than
in the north-west. This is conventional
wisdom; however, rates for the south appear
to be generally much lower than (for
example) the 27 t ha-1yr-1 for the whole of
Spain which was quoted in Europes
environment: The Dobr assessment (EEA, 1995,
p. 155).

Interpretation of the map


General
There are three broad zones of erosion in
pan-Europe: a southern zone, a northern
Erosions impacts across Europe
Zone

Short term (i.e. decades)

Long term (i.e. centuries)

North-west Europe

Off-site: water pollution from agricultural chemicals

On-site: loss of soil productivity

Southern Europe

On-site: mainly loss of soil productivity

On-site: loss of soil productivity

Eastern Europe

On-site: loss of soil productivity


Off-site: water pollution from former industrial
waste, as well as agricultural chemicals

On-site: loss of soil productivity

(10) Iceland was not included in the original study.

Table A3.2

74

Assessment and reporting on soil erosion

Impacts of erosion
The impacts of erosion are not a simple
function of erosion rate. These impacts can
be categorised as on-site and off-site. Offsite problems of water pollution from
agricultural chemicals can result even from
very low rates of soil loss (Harrod, 1994).
Erosions impacts across Europe can be very
generally summarised as described in
Table 2.
Policy implications
EU recognition of the impacts of soil erosion
has to date largely been confined to the
south of Europe e.g. the Corine (11) and
Medalus (12) studies (Stanners and
Bourdeau, 1995). This is principally due to a
focus only on on-site effects. However, if offsite impacts are also considered, then there is
a need for greater EU acknowledgement of
erosion problems elsewhere in Europe. At a
national level, there has been some progress
in this direction.
Along with need for EU recognition of the
Europe-wide nature of erosions impacts is a
need for coordinated scientific endeavour to
tackle it. An immediate need is for an
improved European map of erosion
problems. As far as is possible, this should be
based (for the reasons given above) upon
measured data.
Future work
Soil erosion is a serious problem in Europe,
yet the availability of measured data is very
poor. Thus, effort should be put into:
the establishment of appropriate
monitoring schemes to assess current rates
of erosion (cf. Poesen et al., 1996b; Burt,
1994);
the creation of schemes to bring together
existing measured data, including
information regarding collection
methodologies;
the production of an improved map based
upon these data.
There has some been some recent progress
on the second point. The establishment of
international groups such as the IGBP-GCTE
soil erosion network (Ingram et al., 1996)
and EU COST Action 623 Soil erosion and
global change (see http://

(11) See Corine (1992).


(12) For example, Mairota et al. (1997).

www.cost623.leeds.ac.uk/cost623/) has
enabled erosion researchers to begin to
establish the dialogues which will eventually
lead to a better-harmonised and more freely
available pool of data on erosion. A first
product is GCTE (1997).

Qualitative small-scale soil


degradation assessment databases
The Glasod map
G. W. J. Van Lynden
Project Officer, Soil Degradation and Conservation
International Soil Reference and Information
Centre
In this paper a standardised and
internationally endorsed methodology for
qualitative and mostly small-scale
assessment of soil degradation will be
described. The methodology was first used
for the global assessment of the current status
of human-induced soil degradation (Glasod)
in 1990 and subsequently in a slightly
modified format for other assessments (see
Figures 1 and 4.6 in the main text).
The Glasod map (1990)
Introduction
Recognising the need to obtain a better
overview of the geographical distribution and
the severity of human-induced soil
degradation, the United Nations
environment programme (UNEP)
commissioned the International Soil
Reference and Information Centre (ISRIC)
in 1988 to coordinate a worldwide
programme in cooperation with a large
number of soil scientists throughout the
world to produce, on the basis of incomplete
existing knowledge, a scientifically credible
global assessment of the status of humaninduced soil degradation within the shortest
possible time frame.
Activities included:
the preparation of general guidelines;
subcontracting correlators in 21 regions to
prepare, in close cooperation with national
soil scientists, regional soil degradation
status maps;
correlation of the regional maps into a
world map of soil degradation; and
publishing 5 000 copies of the Glasod map.

Annex III Background papers presented at the workshop

The Glasod map

Publication
The Glasod map was officially released at the
14th International Congress of Soil Science
in Kyoto (August 1990). Subsequently the
map was digitised and a soil degradation
database was created. Thematic maps,
derived from the Glasod database, were
prepared by UNEP/GRID for inclusion in
the World Atlas of Desertification. The
Glasod map and complementary statistics
have been used and cited in numerous
scientific journals and policy documents of
the World Resources Institute, the
International Food Policy Research Institute,
the Food and Agriculture Organisation of the
United Nations, the United Nations
environment programme, and many others.
Scope of the assessment
The assessment was made on a small scale
(1:10 million average) and has a global
coverage. It was based on expert judgments
from national institutions or individual
scientists and addressed the current status of
degradation rather than risk. More than 20
possible soil degradation types were
considered.
Strength and impact of Glasod
Glasod was the first comprehensive soil
degradation overview to be published on a
global scale in a relatively rapid (three years)
and cheap (around USD 300 000) manner. It
raised awareness on soil degradation
problems and created wide interest among
scientists and the general public. It provided
an overview for national and regional
planning and enabled identification of hot
spots for further study. From the received
feedback it was clear that Glasod responded

to a strong apparent need for a global


overview. Multiple requests were received for
national breakdowns or new assessments at
country level. Glasod also showed the need
for an assessment of measures to control
degradation, i.e. showing some good news.
At the same time, the need for an additional
more objective/qualitative approach
(especially for more detailed scales) as well as
the need for data validation and updating
also became obvious.
Who are the users?
Glasod has a wide range of potential and
proven users, such as:
international policy-makers and planners
(e.g. UNEP, FAO, WRI),
national policy-makers and planners,
international conventions and
programmes (CCD, Kyoto Protocol, UNCPB, IGBP),
researchers at national and international
level (NARIs, CGIAR, universities),
education professionals (teachers,
professors, etc.) and students,
environmental organisations (general
public awareness).
Limitations and problems
As a quick and dirty methodology, Glasod
(and its derived successors) also has several
limitations that need to be taken into
consideration. Some of these limitations were
overcome in subsequent assessments.
The small scale makes Glasod less
appropriate for national breakdowns;
The expert judgment approach can lead to
subjectivity;

75

Figure A3.2

76

Assessment and reporting on soil erosion

Figure A3.3

The ASSOD map for south and south-east Asia

Cartographic restrictions at the time of


publication limited the number of
attributes on the map;
The representation of the map items causes
a visual exaggeration: each polygon which
is not 100 % stable shows a degradation
colour, even if only 1 to 5 % of the polygon
is actually affected;
Extent was expressed in classes rather than
percentages;
The map has a complex legend: extent and
degree (severity) are aggregated for four
major degradation types (water and wind
erosion, physical and chemical
deterioration);
Only the dominant main type of
degradation is shown in colour;
Degradation sub-types are only shown by
codes printed in each polygon;
Glasod presented only bad news (doom
scenario).
Follow-up of Glasod / derived initiatives
ASSOD (1997): Assessment of the status of
human-induced soil degradation in south and
south-east Asia
In 1993 an expert consultation of the FAOsupported Asian network on problem soils
recommended the preparation of a south
and south-east Asian soil degradation status
assessment (ASSOD) on a scale of 1:5
million. This study was commissioned by
UNEP to ISRIC and carried out in close
cooperation with FAO and national

institutions in 16 countries. The project used


a modified Glasod methodology, with more
emphasis on the impact of degradation on
productivity and on the rate of degradation
and used a 1:5 million physiographic base
map following criteria outlined in the global
and national soil and terrain digital databases
methodology (SOTER, Van Engelen and
Wen, 1995). The ASSOD project finished in
1997. All information was stored in a digital
database, which is linked to physiographic
units through a GIS. This enables a more
flexible production of outputs: thematic or
regional maps, no restrictions on number of
attributes per polygon, less complex legend.
Soveur (2000): Mapping of soil and terrain
vulnerability in central and eastern Europe
In 1997 FAO and ISRIC initiated the project
on Mapping of soil and terrain vulnerability
in central and eastern Europe (Soveur).
There were three main activities in the
project:
development of a soils and terrain digital
database, at scale of 1:2.5 million, for the
countries under consideration, using the
uniform methodology of SOTER;
assessment of the status of soil degradation,
with special focus on diffuse pollution,
according to a modified Glasod
methodology;
providing the soil geographic and attribute
data for an assessment of the vulnerability
of soils to selected categories of pollutants.

Annex III Background papers presented at the workshop

Implementation
The Soveur project has been implemented in
close collaboration with specialist institutes
from 13 countries: Estonia, Latvia, Lithuania,
Poland, Czech Republic, Slovakia, Hungary,
Romania, Bulgaria, Belarus, Ukraine,
Moldova, and (the European part of) the
Russian Federation. Initial results were
presented and discussed during an
international workshop in October 1999.
Thereafter, the assessment was finalised. In
December 2000 the databases and technical
documentation were released on a CD-ROM
in the FAOs Land and water media series
(No10). This CD-ROM contains information
in the form of databases, maps and reports
on soil, on the soil degradation status and
gives a soil vulnerability assessment for 11
metals in 13 countries in central and eastern
Europe.
Beneficiaries
Target beneficiaries are ministries and
planning bodies in the collaborating
countries who can use the definitive
databases and derived maps for policy
formulation at the national level, for instance
by identifying areas considered most at risk.
The project further contributes to
strengthening the capabilities of national
environmental organisations in central and
eastern Europe, and it can play a significant
role in enhancing scientific cooperation
within Europe on issues of soil degradation
and pollution. Further, it is an integral part of
a global programme on the development of a
world soils and terrain information system, a
world assessment of the status and risk of soil
degradation, and studies of the potential
productivity assessment of the land (cf..
UNCED, 1993).
General degradation guidelines
Based on the experiences with Glasod
(Figure 1), ASSOD (Figure 2) and Soveur
(Figure 3), guidelines for the qualitative
assessment of soil degradation have been
developed that are generally applicable,
scale-independent and offer links to other
standardised methodologies (SOTER,
WOCAT).
WOCAT: since 1992, ongoing
In response to the bad news of Glasod a new
project was initiated to investigate what
measures are being taken to combat
degradation. A consortium of various
national and international organisations,
institutions and individuals, guided by a
management board is undertaking an

The Soveur map for central and eastern Europe

inventory of soil and water conservation


(SWC) worldwide. Through the collection,
analysis and dissemination of existing
experiences, it is expected that mistakes and
duplication of efforts can be minimised.
WOCAT is using a set of comprehensive
questionnaires on technologies, approaches
and mapping respectively that serves as a
framework for the evaluation of soil and
water conservation and a methodology for
data collection at the same time. This
information is stored in an MS-ACCESS
database with a user-friendly menu for
storage, analysis and output of data. Regional
and national training workshops to assist in
the data collection, analysis and output
production have been conducted in over 30
countries, mainly in Africa, Asia and Latin
America, and further SWC evaluation
programmes are ongoing in most of these.
Methodological details
Mapping
As a base map for the assessments in
principle any map with uniquely delineated
polygons can be used, but a (SOTER)
polygon map based on physiography/soils is
recommended. For the Glasod map this was
not yet the case and an IGN world map on a
1:10 million average scale, Mercator
projection with loosely defined
physiographic units was used. The ASSOD
project used a SOTER physiographic map
(1:5 million), while in the Soveur project a
SOTER soil and terrain map on a 1:2.5
million scale served as a basis for the
assessment. For WOCAT several base maps
have been used: the Glasod map for the first
exercises in eastern and southern Africa
(1995), the ASSOD map in Thailand and
China (Fujian province), SOTER maps

77

Figure A3.4

78

Assessment and reporting on soil erosion

Differences between various degradation


assessments
Table A3.3

Comparisons of degradation assessments


Glasod

ASSOD

Coverage

Global

South and south-east Asia


(17 countries)

Central and
eastern Europe
(13 countries)

General

Scale

1:10 million
(average)

1:5

1:2.5

Variable

Base map

Units loosely
defined
(physiography,
land use, etc

Physiography, according
to standard SOTER
methodology

Physiography and
soils according to
standard SOTER
methodology

SOTER maps or
other as
appropriate

Status
assessment

Degree of
degradation +
extent classes
(severity)

Impact on productivity (for


three levels of
management) + extent
percentages

Degree and
impact + extent
percentages

Degree and impact


+ extent
percentages for
major land use
types

Rate of
degradation

Limited data

More importance

As for ASSOD

As for ASSOD

Conservation

No conservation
data

Some conservation data

No conservation
data

No conservation
data, but close link
with WOCAT

Detail

Data not on
country basis

Data available per country

Data available
per country

Depends on scale

Cartographic
possibilities

Maximum two
types per map
unit

More degradation types


defined, no restrictions for
number of types per map
unit

As for ASSOD,
but special
emphasis on
pollution

As for ASSOD

End product

One map
showing four
main types with
severity

Variety of thematic maps


with degree and extent
shown separately

As for ASSOD

As for ASSOD

Database/GIS

Digital
information
derived from
conventional map

Data stored in database


and GIS before map
production

As for ASSOD

As for ASSOD

Source

Individual experts

National institutions

National
institutions

Regional, national
or local institutions

(1:5 million) and larger in other places and


more recently also administrative maps with
districts, provinces, etc. serving as polygons.
The latter has the advantage that many
socioeconomic data are more readily
available for administrative units. Additional
soil and terrain information, as well as other
bio-physical layers such as land cover and
land use, can then be overlaid later on.
Data collection (degradation)
For the degradation assessment the following
data are collected:
major land use type (13)(14) (15),
type of degradation (four main types, about
20 sub-types),
extent of degradation: affected percentage
of (land use area within) polygon,
degree of degradation,
(13)
(14)
(15)

not in GLASOD.
not in ASSOD.
not in SOVEUR .

Soveur

General

impact of degradation,
rate of degradation,
causative factors.
In addition to these, WOCAT collects the
following information on SWC practices:

name of the technology,


SWC category,
extent,
effectiveness,
trend in effectiveness,
period of implementation,
reference to the corresponding technology
questionnaire (describing the same
technology in much higher detail),
productivity trend,
reason for positive or negative productivity
trend.

Annex III Background papers presented at the workshop

Soveur assessment: degree and relative extent of water erosion


Degree and relative extent of water erosion
Wt
COUNTRY

DEGREE
Light

(% of country area)
Moderate

Strong

Extreme

Total

Belarus

1.8 %

6.7 %

0.0 %

0.0 %

8.5 %

Bulgaria

21.0 %

16.6 %

2.1 %

0.0 %

39.8 %

Czech

8.9 %

5.7 %

0.5 %

0.0 %

15.1 %

Estonia

0.8 %

2.4 %

0.0 %

0.0 %

3.2 %

Hungary

2.6 %

8.5 %

10.1 %

0.0 %

21.2 %

Latvia

0.0 %

11.3 %

0.1 %

0.0 %

11.4 %

Lithuania

3.0 %

7.1 %

0.3 %

0.0 %

10.4 %

Moldova

0.3 %

6.9 %

27.6 %

0.0 %

34.8 %

Poland

0.0 %

6.4 %

0.4 %

0.0 %

6.7 %

Romania

11.0 %

7.2 %

0.0 %

0.0 %

18.2 %

Russia

0.0 %

0.0 %

4.2 %

0.0 %

4.2 %

Slovakia

0.5 %

0.8 %

4.1 %

0.0 %

5.4 %

Ukraine

2.9 %

12.1 %

0.4 %

0.0 %

15.4 %

Wd
COUNTRY

DEGREE
Moderate

Strong

Extreme

Total

Belarus

0.1 %

0.0 %

0.0 %

0.0 %

0.2 %

Moldova

0.3 %

0.4 %

0.3 %

0.1 %

1.0 %

Poland

0.0 %

2.2 %

0.3 %

0.0 %

2.5 %

Romania

0.2 %

1.7 %

6.7 %

0.0 %

8.6 %

Slovakia

0.3 %

2.4 %

4.1 %

0.0 %

6.8 %

Ukraine

0.0 %

0.8 %

1.8 %

0.0 %

2.5 %

Moderate

Strong

Extreme

Total

Wo
COUNTRY

Light

(% of country area)

DEGREE
Light

(% of country area)

Lithuania

0.0 %

0.4 %

0.0 %

0.0 %

0.4 %

Romania

2.5 %

1.5 %

0.0 %

0.0 %

4.0 %

Ukraine

0.2 %

1.3 %

0.9 %

0.0 %

2.5 %

Note: Wt = water erosion (terrain deformation)


Wd = water erosion (loss of topsoil)
Wo = water erosion (off-site effects)

Results of the assessment


Output production (degradation)
Whereas for Glasod, due to conventional
preparation of the map, only a hard copy
world map was produced (see Figure 1), the
ASSOD and Soveur data can be presented in
a much more flexible way: various digital and
hard copy thematic maps (Figures 2 and 3)
and/or regional selections; tables (Table 2)
and graphs on area coverage for all or
selected types, degree of degradation,
impact, rate, etc.
Future developments?
More regional and national qualitative
assessments
More specific and quantitative assessments
Glasod revisited (new global assessment)

Links to action plans for remediation

Indicators of soil erosion at the


ETC/Soil
O. Dwel and J. Utermann
Bundesanstalt fr Geowissenschaften und Rohstoff
(BGR)
Introduction
The European Environment Agencys (EEA)
main mission is to support sustainable
development through the provision of
relevant, reliable, targeted and timely
information to policy-makers and the general
public. For this purpose the EEA is
establishing a monitoring and reporting
system based on indicators. In the period

79

Table A3.4

80

Assessment and reporting on soil erosion

Figure A3.5
Source: Agricultural
statistics (Eurostat in
report of DG VI); Indicator
fact sheet AG6
Agricultural
environmental efficiency;
New Chronos: Data set
name: Number and area
of agricultural holdings
including mountains

Intensification of agricultural productivity as a driving force (pressure) indicator Intensification of


agricultural productivity measured in terms of annual work units (AWUs) as compared to gross value added
(GVA), EU-15, 198397
Index
120

100

80

60

40

20

0
1983

1985

1987

1989

1991

1993

1995

1997

Years
AWU index (1983 = 100 = 10 784.3 units)
GVA index (1985 = 100 = 141 265 ECU)
ha/holding index (1990 = 100)

Note: Additional information about intensification can be derived from the increase of total agricultural area per
agricultural holding.

199699 the European Topic Centre on Soil


(ETC/S) supported the EEA work
programme providing information and
contributing to EEA reporting on
environmental issues related to soil.
The main causes of soil loss and
deterioration in Europe are considered to be
soil sealing, soil erosion and local and diffuse
contamination (EEA, 1999a). The following
paper deals with the physical soil degradation
patterns of soil erosion and soil sealing.
The indicator concept
In 1993, the OECD presented a core set of
indicators for environmental performance
reviews (OECD, 1993). Indicators were
defined as a parameter, or a value derived
from parameters, which points to/provides
information about/describes the state of a
phenomenon/environment/area with a
significance extending beyond that directly
associated with a parameter value (OECD,
1993).
The basic OECD criteria for indicator
selection have been applied by the EEA. The
criteria can be summarised under three
headings: policy relevance and utility for
users, analytical soundness, and measurability
(see Part I, Section 2.1 of the main text for a
full list). When developing indicators these
criteria should be taken into account.

Different human activities (driving forces)


exert pressures on the environment and
change its quality (state). The change of
environmental conditions has impacts on
other environmental issues. Society responds
to the changes and impacts through
environmental, general economic and
sectoral policies. Taking this chain of causes
and effects into account the EEA has
developed the DPSIR framework (driving
forces, pressures, state, impact, responses)
(cf. Gentile, 1999a). The development of
relevant indicators for reporting makes use of
this framework.
DPSIR applied to soil erosion
In a first attempt the DPSIR assessment
framework has been applied to the soil
degradation patterns soil erosion and soil
sealing. The aim was to develop short-term
indicators for actual reporting as well as a
long-term approach for a periodical
reporting system.
Figure 2.4 in the main text presents the
DPSIR assessment framework applied to soil
erosion. The intensification of agriculture
increases the risk of insufficiently sustainable
land use practices. Unsustainable land use
practices in themselves enhance the risk of
soil degradation, e.g. physical soil
deterioration.

Annex III Background papers presented at the workshop

Indicator-based approach for monitoring soil erosion

Soil loss = f (R, K, L, S, C, P)


Potential soil
erosion risk

Data
needs

Soil (K)
Relief (S, L)
(climate (R))

Suitable
scale:

Small scale
 1:1 000 000

Time
variability:

Actual soil
erosion

Actual soil erosion risk

land use (C)


Principal land use
(arable land/pastures/
forests/others)

Actual land use of arable


land (maiz, sugar beets,
potatoes, summer grains)

Medium scale
1:25 0001:100 000

Large scale
< 10 000

Increasing time variability


no

high

Indicatorrequirements:
Theoretic founded
Measurability
Trends over time
Data availability

Control practices (P)


esp.: ground cover by
plant residues or
vegetation

+
+
+

+
+
-(+)
+

Possible driving forces and/or pressure


indicators should aim to describe the
intensification of agriculture and its increase.
Figure 1 gives an example: the gross value
added (GVA) reflects the efficiency of the
agricultural production. The data on working
hours, expressed as annual working units
(AWUs) give information on the actual
volume of labour devoted to farming.
An increasing GVA, produced by a
decreasing number of people employed in
the agricultural sector, shows the trend of
intensification of the agricultural
production. This trend is confirmed by the
trend towards larger units of agricultural
holdings. The indicator has to be seen in
combination with other indicators describing
the agricultural section, e.g. the trends of
fertiliser and pesticide uses.
Short-term information about the current
state of soil erosion is given in the report
Environment in the European Union at the turn of
the century (EEA, 1999a). The soil chapter of
this report contains two tables (Tables 3.6.2
and 3.6.3) dealing with the area affected by
water erosion and the total amounts of soil
losses due to water erosion. The data are
derived from questionnaires from statistical
institutions (e.g. Eurostat) and the EEA.
Since in most European countries there are
no data available on soil erosion, these tables
are rather fragmentary they only contain
information from Germany, Spain, Austria
and Iceland.

+
+
+
+(-)

+
+
+
-

An attempt at a data update by the ETC/S in


1999 resulted only in few additional data and
in some cases (Germany) the data previously
available were withdrawn.
A long-term approach should focus on the
abovementioned basic criteria for indicator
selection. In order to cover all European
countries, a unique methodology has to be
implemented. The methodology has to
enable a periodical monitoring to identify
trends and effects of response measures.
Soil losses due to water erosion are a function
of climate, soil, relief, vegetation and
protection measures (cf. for example the
universal soil loss equation (USLE)
(Wischmeier and Smith, 1978).
One problem is the high variability of the
data needed, both in terms of scale and time.
But if it were possible to make available data
related to control practices, e.g. mulching or
other measures increasing the ground cover
(by vegetation, plant residues or others), this
would be an indicator for soil erosion that is
theory based, measurable and able to show
trends over time. Figure 2 summarises the
proposed approach.
Following this approach the first step should
be to identify areas with a high potential of
soil erosion risk with regard to soil erodibility
and relief. The second step is to focus on
areas where there is a high risk of actual soil
erosion taking into account the principal
form of land use (e.g. arable land).

81

Figure A3.6

82

Assessment and reporting on soil erosion

Table A3.5

Calculated annual gross erosion rates on the basis of the sediment delivery ratio (selected data: minimum
and maximum annual gross erosion per country)

Source: Data collection


request, 1999 (EEA-ETC/
S).
Finland

River name

Sampling
periods

Kokemenjoki

9196

CA
(catchment
area) (km2)
27 046

Suspended
solids
(t/a)

SDR

Annual
sediment
loads/CA
(t/ha*a)

91 463

0.03

0.03

Annual gross
erosion/CA
(t/ha*a)
1.1

Iijoki

9196

14 191

17 192

0.04

0.01

0.3

FYROM
(Macedonia)

Vardar

7186

21 350

2 207 520

0.03

1.0

31.9

Sataeska

7890

351

32 483

0.1

1.0

9.3

Germany

Oder

92/97

112 950

441 000

0.02

0.04

2.2

Donau

92/97

77 053

2 557 000

0.02

0.3

16.1

As a third step the ground cover should be


monitored for example, allowing an
assessment of the state of soil erosion by
water. In order to be able to translate the
monitored ground cover data into actual soil
losses a fourth step is needed, namely the
monitoring of real actual soil erosion losses
in selected test areas.
Especially with regard to agricultural land
this approach means: the higher the share of
crops which increase the risk of soil erosion
(row crops, e.g. corn, sugar beet, potatoes)
of total arable land in areas with a high
potential soil erosion risk, the higher the
actual soil losses due to soil erosion, unless
accompanying protection measures are
applied.
Soil, climatic and relief conditions cannot be
changed by human activities, at least not in
the short term. Hence ground cover
measures could be used to combat soil
erosion in other forms of land use as well.
Another indicator approach towards
assessing the situation of soil erosion in
Europe is based on the measurement of
sediment loads in rivers and streams. The
idea was to use the sediment delivery ratio for
these estimates. The sediment delivery ratio
is the fraction of the gross erosion that is
expected to be delivered to the point of
drainage area under consideration (Mitchell
and Bubenzer, 1980). On the basis of an
equation published by Wischmeier (16)
(1975, 1976) an attempt was made to
calculate annual gross erosion rates (t/ha*a).
Table 1 gives selected results (minimum and
maximum calculated annual gross erosion
per country). The data have been obtained
through the data collection request of the
ETC/S this year. Data were delivered by
Finland, Macedonia and Germany.

Organised in classes of < 1 t/ha*a (none to


low erosion), 110 t/ha*a (moderate
erosion), 1020 (high erosion) and (10)30
t/ha*a, the results show almost none to low
erosion rates in Finland, moderate (northern
part) to high rates (southern part) in
Germany, and up to very high erosion rates
in the southern European area of
Macedonia.
The present approach (which is just at a
conceptual level) will have to be discussed in
depth. Among other things, one has to take
into account that originally the sediment
delivery ratio was used to determine largescale water pollution from agricultural land
(< 1:5 000).
Nevertheless data on suspended solids give
information about the impact of soil erosion
on other media and should at least as an
impact indicator be integrated into an
environmental monitoring network.
Summary
Table 2 summarises the approaches discussed
in this paper to determine state indicators,
both short term and long term for the soil
degradation patterns of soil erosion.
Short-term state indicators for soil erosion
are presented in the soil chapter of the Turn
of the century report published by the EEA in
1999 (EEA, 1999a). They describe the area
affected by erosion and total amounts of soil
losses due to erosion in selected European
countries based on questionnaires and data
collections by different institutions (OECD,
Eurostat, EEA-ETC/S). Data are only
fragmentary (only few countries available),
and data reliability remains to be checked.

(16) SDR = 0.02 + 0.385 * CA 0.2; SDR = sediment delivery ratio, CA = catchment area.

Annex III Background papers presented at the workshop

Indicator approaches for soil erosion: State of the art

Short-term approach

State
indicators

Data
sources

Area affected by erosion


Soil losses due to
erosion

Long-term approach

ground cover from


vegetation and other
protection measures (e.g.
mulching) in areas of high
potential soil erosion risk

To handle the soil erosion problem in the


long run, a special approach is proposed,
which is based on a map of potential soil
erosion risk combined with periodical
monitoring by means of remote sensing
validated by ground monitoring in selected
test areas. Especially with regard to
agricultural land this approach means: the
higher the share of crops which increase the
risk of soil erosion (row crops, e.g. corn,
sugar beet, potatoes) of total arable land in
areas with a high potential soil erosion risk,
the higher the actual soil losses due to soil
erosion, unless accompanying protection
measures are applied.
There are data needs concerning the ground
cover in areas of high erosion risk. A
European-wide map of potential soil erosion
risk and data of ground cover due to
vegetation and other protection measures
(e.g. mulching) would reveal data gaps. One
way of obtaining additional data would be to
cooperate with other European institutions,
e.g. the Space Application Institute (JRCSAI).
Apart from agricultural land use, ground
cover is a possible indicator for other land
uses potential leading to soil erosion.

GISCO databases and tools to derive


driving force/pressure indicators for
soil erosion
Paul Campling
Ground for GIS, Katholieke Universiteit Leuven
Costas Kosmas
Agricultural University of Athens
This paper provides an overview of the
GISCO databases and tools which are
currently available for providing information
on indicators for soil erosion in general and
driving force and pressure indicators for soil

Data
availability

Data
reliability

Different statistical
institutions
Questionnaires and data
collections

(!)

Map of potential soil


erosion risk
Periodical monitoring by
remote sensing
combined with ground
validations in test areas

! (?)

! (?)

erosion in particular. The paper is broken


down into five sections:
Introduction to GISCO databases and tools
Overview of driving force/pressure
indicators proposed by the EEA
GISCO and driving force/pressure
indicators
Proposed indicator framework model
Remarks and conclusions
Introduction to GISCO databases and tools
The geographic information system of the
European Commission (GISCO) databases
and tools are developed, maintained,
updated and distributed by Eurostat, located
in Luxembourg. GISCO is a multi-layered,
multi-scale geographic reference database
containing topographic and thematic layers.
The system is developed so that it can be used
as a practical tool for policy-makers, who are
able to view the spatial component of data
sets, bring in supplementary data, and carry
out spatial analysis and overlay functions.
There are five different scale layers: 1:25
million, 1:10 million, 1:3 million, 1:1 million,
and 1:100 000, of which the 1:1 million is
most common. Tools have been developed
for standardised cartographic production,
which can be customised for advanced spatial
analysis functions. The ArcView mapping
tool has a series of Avenue scripts developed
for inexperienced ArcView users. The
ArcInfo mapping tools represent a suite of
AML programs that are suitable for use by
experienced ArcInfo users
The GISCO themes that are related to soil
erosion are: environment, hydrography, land
resources, altimetry, administrative
boundaries, infra-regional statistics, and
community support frameworks.
Environment
Potential and actual soil erosion risk
databases are available for southern Europe.
Actual soil erosion risk maps are derived by

83

Table A3.6

84

Assessment and reporting on soil erosion

combining four sets of factors: soil erodibility,


rain erosivity, slope angle and vegetation
cover. Soil erodibility is assigned a value
between 1 (low) and 3 (high), on the basis of
soil physical properties. Rain erosivity is
assigned a value between 1 (low) and 3
(high), based on the Fournier precipitation
index and the Bagnouls-Gaussen aridity
index. Slope angle is assigned a value
between 1 (very gentle to flat) and 4 (very
steep), on the basis of a variety of
topographic sources (topographic maps,
digital elevation models and satellite
imagery). Vegetation cover is assigned a value
of 1 (fully protected) or 2 (not fully
protected) from the Corine land cover
database (Figure 1). This represents a
simplified universal soil loss equation
(USLE) approach and results in a map at 1:3
million scale. The potential soil erosion risk
is derived by excluding the vegetation cover.
The actual soil erosion risk map available at
GISCO is only available for Portugal, because
at the time the soil erosion maps were
created only the Corine land cover (LC) map
of Portugal was available.
Land quality maps are also only available for
southern Europe, and combine factors in a
similar manner to the potential and actual
soil erosion risk maps. Land quality maps
combine four sets of factors: soil, climate,
slopes and land improvements, on a 1:3
million scale.

different countries, or because of the short or


irregular periods for which stations have
been maintained.
The climate interpolated database is made
on 50 x 50 km grid cells covering Europe and
Magreb. The monthly data have been
recalculated from long-term average daily
data for the period 197599 for the following
parameters: absolute minimum temperature;
average minimum temperature; absolute
maximum temperature; sum of precipitation;
sum of potential evaporation; and sum of
global radiation.
The land cover database is derived from the
Corine land cover for the year 1990, and is
distributed as grids of 100 m and 250 m
resolution. The minimum mapable unit for
land cover is 25 ha, being based on visual
interpretation of Landsat and SPOT
multispectral data. There are three levels of
classification, with the third level having 44
land cover classes.
The soils database (version 2) from the
European Soil Bureau (JRC Ispra) is
currently available at GISCO, based on FAO
nomenclature and on a 1:1 million scale.
Attributes include soil mapping units
(SMUs), polygons of the same soil type, and
soil typological units (STUs), indicating the
main soil types contained in SMUs. A more
recent version is held at the European Soil
Bureau (JRC Ispra), with a wider European
coverage, and more detailed soil physical
attributes.

Hydrography
The hydrography databases include rivers
and lakes coverages on a 1:3 million scale
and catchment boundaries on a 1:3 million
scale. JRC (ISPRA) is currently preparing a
catchment boundaries database on a 1:1
million scale for inclusion in the GISCO
databases in the near future.

Altimetry
The digital elevation model is a panEuropean raster coverage providing
elevation heights for 1 x 1 km grid cells on a
1:3 million scale.

Land resources
Climate data are provided from 5 308 stations
in the EU (12 Member States). The two main
climatic variables are precipitation (average,
maximum 24-hour rainfall, number of rain
days, average snowfall, number of snowfall
and snow cover days) and temperature
(average, maximum, minimum, absolute
monthly maximum and minimum, number
of frost days). Other climate attributes
include relative humidity, vapour pressure,
atmospheric pressure, bright sunshine,
evapotranspiration, wind speed and cloud
cover. There are more gaps in these records
because of inconsistencies in the definitions
and measurement procedures used in

Administrative boundaries
The nomenclature of territorial units for
statistics (NUTS) regions serve as a base map
of regional boundaries covering the entire
EU territory. The NUTS nomenclature
subdivides the EU economic territory into six
administrative levels, from country (level 0),
through regional (levels 1, 2, 3) to local
(levels 4, 5) level. At present, three NUTS
versions (V5, V6 and V7) for three scale
ranges (1 million, 3 million and 10 million)
are maintained at GISCO. The NUTS
coverages provide the means to spatially
present agricultural statistical survey and
census data from the farm structure survey
and Structural Fund databases. The

Annex III Background papers presented at the workshop

boundary coverages delineate the regions


while the point coverages provide a label for
each region. Associated tables contain basic
information such as the regions name and
area.
Infra-regional statistics
The degree of urbanisation is derived from
population census data from 1981 and 1991
on a 1:1 million scale. The coverage has three
density classes: densely populated area
(contiguous set of local areas with a
population density greater than 500
inhabitants per km2, and a minimum total
population of 50 000 inhabitants);
intermediate area (contiguous set of local
areas, not belonging to the densely
populated area, with a population density
greater than 100 inhabitants per km2, and a
minimum total population of 50 000
inhabitants or adjacent to a densely
populated area); and thinly populated area
(contiguous set of local areas belonging
neither to a densely populated nor to an
intermediate area).The local area
corresponds to the NUTS 5 administrative
units of the Member States. Changes in the
degree of urbanisation can indicate trends in
the degree of soil sealing.
Community support
Databases found under the community
support theme provide quasi socioeconomic
information on regions that receive support
from the EU.
The less-favoured areas, originally created at
the Directorate-General for Agriculture, are
part of the Structural Funds programme,
which represents areas defined as regions
where economic activities, from the
agricultural point of view, are difficult to
pursue. The criteria, developed though
consultation with the Member States, include
mountainous regions, areas in danger of depopulation, and areas with specific handicaps
(for example, desertification, marsh lands,
salinisation). The coverage is provided on a
1: 3 million scale.
The Structural Funds is made up of five data
sets that indicate the areas of the EU eligible
for support from Structural Funds in the
following periods: 198993; 199499 (12
Member States); 199499 (15 Member
States); 199499 (15 Member States, update
reference year 1997); and 200006 (15
Member States). The coverage is provided on
a 1:1 million scale.

Overview of driving force/pressure


indicators proposed by the EEA
Within the DPSIR assessment framework the
EEA has provided a list of driving force (D)
and pressure (P) indicators for soil in general
(EEA, 2000), and soil erosion in particular
(EEA-ETC/S, 1999). Agricultural
intensification is identified as both a driving
force (for soil erosion in particular) and a
pressure indicator (for soil in general)
(Figures 2.2 and 2.4 in the main text). Under
agricultural intensification (or the degree of
agricultural land use), the following
characteristics can be identified: fertiliser use
and trend (P); farm size and trend (D/P);
field size and trend (D/P); crop yield and
trend (D/P); and stocking rate and trend
(P). Forest fires and deforestation are
identified as pressure indicators for soil
erosion.
GISCO and driving force/pressure
indicators
GISCO databases, such as NUTS and Corine
land cover, can be used together with farm
structure survey (FSS) data from Eurostat to
derive indicators of agricultural
intensification (Figure 3). Farm structure
survey data include information of crop type
and area, crop yields, and livestock type and
number at the NUTS 2 and 3 level. NUTS 2
data are collected on a bi-annual basis and
are based on representative surveys, whereas
the NUTS 3 data are census information
collected on a decade basis. Therefore trends
in crop yields and livestock stocking rates can
be detected at the NUTS 2 level.
Correspondence tables that relate crop types
to Corine land cover types can be used to
spatialise or disaggregate the agricultural
statistical data. Both Eurostat and the JRC are
currently conducting research in the domain
of relating farm structure survey data to the
Corine land cover data set.
Proposed indicator framework model
Agricultural intensification should not be
regarded as the sole driving force indicator
for soil erosion, as there is not a proven link
between agricultural intensification and land
degradation. Intensive farming often
encourages good conservation practices,
because land becomes scarce and increases
in value, and therefore farmers adopt, by
necessity, soil nurturing practices. Other
driving forces may be just as, or even more,
important. Anthropogenic changes in
agricultural practices, urbanisation and
tourism may cause changes in land use and
vegetation. In addition, increase in the

85

86

Assessment and reporting on soil erosion

frequency and area of forest fires may cause


changes in the natural vegetation. Decrease
in the local human population may lead to
the abandonment of terrace maintenance,
because of insufficient labour, whereas a
sudden increase in population, due to
migration, may put vulnerable soils under
greater risk of erosion. Climate indicators are
also important factors, such as in
Mediterranean areas where 300 mm rainfall
is seen as the threshold annual rainfall
amount with areas receiving less than 300
mm rainfall per year being better kept as
natural vegetation and areas receiving more
than 300 mm rainfall being suitable for rainfed agriculture.
Land cover changes and precipitation
indicators, on the other hand, are seen as
better pressure indicators for soil erosion, as
these directly influence the degree of soil
erosion. Land cover changes, including
forest fires and deforestation, can be
detected by combining the reference land
cover database, Corine land cover, with
vegetation changes indices from NOAAAVHRR and SPOT Vegetation. Precipitation
regimes can be detected using the GISCO
climate coverages and the monitoring
agriculture by remote sensing (MARS)
meteorological database.
Remarks and conclusions
GISCO databases provide EU-wide georeferenced bio-physical and socioeconomic
information, generally on a scale of 1:1 000
000. An actual and potential soil erosion risk
map based on a simplified USLE approach is
available. Agricultural intensification can be
assessed by combining farm structure survey
and Corine LC data. But agricultural
intensification trends are only possible at
NUTS 2 level, which may be regarded as
being too coarse for application to soil
erosion issues. There is currently ongoing
research at the JRC and Eurostat in regard to
combining agricultural statistical data
collected at administrative level and raster
information of land cover provided by the
Corine LC. It is argued that agriculture
intensification, in isolation, could be
misleading and therefore it is proposed to
include human population, land
development, tourism, transport, natural
events and climate change with agricultural
intensification for driving force indicators of
soil erosion. Land cover change and
precipitation can be used for pressure
indicators of soil erosion, as they are seen to
be directly influencing the degree of soil

erosion. Land cover change can be detected


by combining the Corine LC with vegetation
change monitoring techniques (AVHRR/
Vegetation). Precipitation can be derived
from the GISCO climate and/or the MARS
meteorological databases.

Regional assessment of the impact of


soil erosion by water
Anne Gobin Laboratory for Experimental
Geomorphology Katholieke Universiteit Leuven
Mike Kirkby School of Geography University of
Leeds
Soil erosion indicators of state
Soil erosion is a natural process, occurring
over geological time. Most concerns about
erosion are mostly caused by water and are
related to accelerated erosion, where the
natural rate has been significantly increased
by human activities that cause changes in
land cover and management.
In Europe, soil erosion is caused mainly by
water and, to a lesser extent, by wind. In the
Mediterranean region, water erosion results
from intense seasonal rainfall on often fragile
soils located on steep slopes. The area
affected by erosion in northern Europe is
more restricted and moderate rates of water
erosion result from less intense rainfalls
falling on saturated, easily erodible soils.
However, these findings are based on
fragmentised and non-standardised
information.
Soil erosion is widely recognised to be patchy
both in time and in space. A major event may
occur in one place but leave the adjacent
field or plot untouched. In addition the lack
of widespread soil loss measurements
hamper effective interpolation between the
limited sites available. Soil loss measurements
or observations typically stretch over a period
of three to five years, and make temporal
extrapolations difficult. The lack of data and
the patchy nature of soil erosion make the
development of an indicator of state a
difficult process.
Ultimately, it is the area that is affected by soil
erosion and an estimate of the expected
magnitude in a particular area that policymakers need to know in order to formulate a
sound soil protection policy. Regional soil
erosion assessment is therefore needed on a
European scale in order to make objective
comparisons that may provide a basis for

Annex III Background papers presented at the workshop

further environmental analysis, economic


statements or policy development. This
paper deals with methods to present soil
erosion on a regional scale.
The revised DPSIR assessment framework
The result of the application of the DPSIR
and MF/MI assessment tools to soil erosion is
the identification of a set of policy-relevant
indicators. However, it has to be recognised
that there is a huge difference between actual
soil erosion and soil erosion risk, which is not
adequately reflected in the present
framework (EEA-ETC/S, 1999). Indicators
describing the driving forces and pressures
may affect the risk of soil erosion, but they
may not affect soil erosion in itself. A
mechanism is therefore needed to jointly
estimate the actual erosion and the risk,
based on links between the identified driving
force and pressure indicators, and based on
an estimation or measurement of what is
actually happening. It is proposed to modify
the current DPSIR assessment framework
(EEA-ETC/S, 1999; EEA, 2000) to include
some of the considerations in a revised
DPSIR scheme, presented in Figure 2.5 in the
main text. Agricultural intensification is seen
as the most important driving force.
However, tourism and transport could be
added to the list. The effect that driving
forces have in common is that they change
the land cover, which is the major pressure
indicator for soil erosion.
Processes of soil erosion by water
Slope sediment transport processes consist of
weathering followed by transport of the
regolith. For both weathering and transport,
the processes can conveniently be
distinguished as chemical, physical,
biological and anthropogenic. Most slope
processes are assisted by the presence of
water, which helps chemical reactions, makes
masses slide more easily, carries debris as it
flows and supports the growth of plants and
animals.
Material may be detached by raindrop impact
and flow traction, and transported either by
saltation through the air or by overland water
flow. Combinations of these detachment and
transport processes give rise to the three
main processes, rainsplash, rainwash and
rillwash.
Runoff is the most important direct driver of
severe soil erosion. Processes that influence
runoff must therefore play an important role
in any analysis of soil erosion intensity, and

measures that reduce runoff are critical to


effective soil conservation.
Regional assessment methods of soil erosion
Methods
Regional assessment methods of soil erosion
include distributed point data, factor or
indicator mapping and process modelling.
All of these methods require calibration and
validation, although the type of validation
needed is different for each category. There
are also differences in the extent to which the
assessment methods identify past erosion and
an already degraded soil resource, as
opposed to risks of future erosion, under
either present climate or land use, or under
scenarios of global change.
One important form of erosion assessment is
from direct measurements or field
observations of erosion features and soil
profile truncation. The main advantage of
measurements is that they are unambiguous
where they exist, and give a good indication
of the current state of degradation of the soil
resource. The main disadvantage of field
observations is that they provide little or no
information about when erosion occurred,
which has little bearing on current or
prospective erosion hazards. The distributed
point data method requires validation to
standardise differences in measurements, in
the intensity of study of different areas and in
the clarity of suitable features on different
soil types. On their own these methods
cannot provide a complete picture except for
small sample areas, and require the use of
other methods to interpolate between areas.
Since many of the processes and factors that
influence the rate of erosion are well known,
it is possible to rank individual factors for
susceptibility to erosion, providing a series of
erosion indicators. Individual indicators may
be mapped separately, but it is more
problematic to combine the factors into a
single scale, by adding or multiplying suitably
weighted indicators for each individual
factor. There are difficulties both about the
individual weightings and about the assumed
linearity and statistical independence of the
separate factors. The method should
therefore be most effective for identifying the
extremes of high and low erosion, but less
satisfactory in identifying the gradation
between the extremes. Despite these
theoretical limitations, factor or indicator
mapping has the considerable advantage that
it can be widely applied using Europe-wide

87

88

Assessment and reporting on soil erosion

geographic data on topography, soils, climate


and land cover/use.
The third method for Europe-wide soil
erosion assessment is the application of a
process model. Process models have the
potential to respond explicitly and rationally
to changes in climate or land use, and so
have great promise for developing scenarios
of change and what-if analyses of policy or
economic options. Although it is the most
generally applicable and replicable method,
the challenge is to relate coarse scale
forecasts to available erosion rate data, much
of which is for small erosion plots and
catchments. Nevertheless this method has
the potential to provide a rational physical
basis to combine factors which can be
derived from coarse scale GIS, and overcome
the difficulties about weighting and intercorrelation which are encountered in purely
factor-based assessments. Set against the
advantages, process models can only
incorporate the impact of past erosion where
this is recorded in other data, such as soil
databases. Models also generally simplify the
set of processes operating, so that they may
not be appropriate under particular local
circumstances.
A number of approaches that present a
regional assessment of soil erosion by water
have been reviewed: Glasod, hot spots, RIVM
and Corine. Both the Glasod and hot-spot
maps were classified as methods based on
distributed point data; the RIVM and Corine
maps were classified as factor- or indicatorbased maps.
The Corine soil erosion risk maps are the
result of an overlay analysis of factors based
on topography, soils, precipitation and land
cover maps, enabling the evaluation of the
soil erosion risk category. The methodology
is based on a considerable simplification of
the universal soil loss equation (USLE), a
regression based model, for which there is a
massive database for US conditions, but little
systematic data for Europe. Potential soil
erosion risk is defined as the inherent risk of
erosion, irrespective of current land use or
vegetation cover, and represents the worst
possible situation. Actual soil erosion risk
reflects the protective influence provided by
present land cover, and the dangers inherent
in changes in land use practices. The Corine
assessment is restricted to southern Europe,
whereas present needs for erosion data apply
to the whole of Europe.

The RIVM approach combines a baseline


assessment of erosion risk with climate and
economic projections to generate scenario
projections for 2010 and 2050. RIVM makes
explicit use of scenario projections, a feature
lacking in other approaches, but the map at
50 km resolution cannot readily be
interpreted at sub-national scales. The
approach also has the advantage of
combining physical and economic elements
within a single framework. The erosion
impact module generates a water erosion risk
index based on three main parameters:
terrain erodibility, rainfall erosivity, and land
use pressure. The RIVM soil erosion model is
a factor model, like Corine, but, although
initiated six to eight years later, is in many
ways a still more simplified approximation to
the imperfect USLE model. Neither the 50
km resolution nor the implementation of the
factors contributing to erosion is seen as
providing a state-of-the-art assessment.
The purpose of the hot-spot study was to
support the joint message on the need for a
pan-European policy on soil, identifying hot
spots of degradation in Europe and
examining environmental impacts leading to
change and particularly degradation of soil
functions. The work involved compilation of
data available and incorporation into a GIS.
The data provides general or particular
information about water erosion for
approximately 60 sites or small regions across
Europe, with measured erosion rates, which
could be placed on the map at 35 sites
classified in three broad zones. The sporadic
distribution and episodic occurrence of soil
erosion and the uneven density and quality of
local measurements makes the data set very
ill-suited to effective interpolation. However,
the data set could be very meaningfully used
in combination with factorial or process
modelling methods.
The main objective of the global assessment
of soil degradation (Glasod) project was to
strengthen the awareness of decision-makers
on the risks resulting from inappropriate
land and soil management to the global wellbeing. Glasod is based on responses to a
questionnaire that was sent to recognised
experts in all countries. The dependence on
expert judgments provides very little control
or objectivity in comparing the standards
applied by different experts for different
areas. An updated version of the European
part of the (Glasod) map was made on the
basis of completed and returned
questionnaires. The degree of detail of the

Annex III Background papers presented at the workshop

Pesera project scientific structure

Phase 1: Pesera model development


Modelling strategy (WP1):
Sediment transport equation
Basis for including other
erosion types

Spatial and temporal


resolution linkages (WP2):
Selection of test areas
Climate generator
Aggregation techniques
Error analysis

Calibration (WP3):
Selection of sites
Calibration with measured
soil erosion rates
Database on erosion rates

Model code (WP1):


in AML (ArcInfo) and/or C++

Phase 2: Pesera model testing


Soil cover module (WP4&5):
SPOT Vegetation image
processing
Soil cover algorithm

Validation on a European
scale (WP5):
Transfer functions and
interpolation algorithms
Comparison with Corine
Database development

Validation at low
resolutions (WP4):
Country case studies
Transfer functions and
interpolation algorithms
Comparison with expert
systems, Corine, USLE
Database development

Validation at high
resolutions (WP3):
Selection of catchments
Transfer functions and
interpolation algorithms
Comparison with USLE,
Eurosem
Database development

Phase 3: Pesera model application

Application on a country and


European scale (WP4&5):
Maps and reports on current soil
erosion rates

Scenario analysis (WP6):


Climate change
Land use change

User interface (WP7):


GUI application

User groups (WP7):


Workshops with end-users
Workshops with expert users

information received varies greatly. It must


also be noted that the scale of the maps (1:10
000 000) limits the detail that can be shown,
providing a minimum resolution of
approximately 10 km. Despite its limitations,
it is the only approach which has been
applied on a worldwide scale.

to the more complex USDA WEPP model.


The pan-European soil erosion risk
assessment method is presented in the next
section as a new method to forecast soil
erosion based on a process model in
combination with validation against field
measurements.

Although the USLE has been the most widely


applied model in Europe, it is now widely
considered to be conceptually flawed, and
other models are now emerging, based on
runoff thresholds or the MIR (minimum
information requirement) approach applied

Process modelling to assess regional soil


erosion: Pesera
Process modelling methods allow for a more
quantitative forecast, which is important as a
critical control on soil erosion. Current
models are designed to assess soil erosion at

89

Figure A3.7

90

Assessment and reporting on soil erosion

Figure A3.8

The Pesera model structure

Rainfall
intensity
Crown cover
distribution
Infiltration
into the soil
At-a-point
runoff
generation

Duration of
intense
showers

Variation of
infiltration
downslope
Accumulation
of runoff
downslope

high resolutions, and are not suitable to


develop regional soil conservation strategies.
The pan-European soil erosion risk
assessment project (Pesera), an EU fifth
framework project (Gobin et al., 1999), has
developed a physically based and spatially
distributed model to quantify soil erosion in
a nested strategy of focusing on
environmentally sensitive areas relevant on a
European scale (Gobin and Govers, 2002).
The objectives of Pesera are threefold and
link to three well-identified project phases
(Figure A3.7).
Project phase 1 focuses on the development
of a process-based and spatially distributed
model to quantify soil erosion and assess its
risk across Europe. The model is intended as
a regional diagnostic tool, replacing
comparable existing methods, such as the
universal soil loss equation, which lack a
sound physical basis and compatibility with
higher resolution models. The model will be
calibrated and validated with existing
information on soil erosion rate
measurements and a prediction error will be
attached to model outputs. This will entail
the development of a modelling strategy,
sensitivity analysis, temporal and spatial
aggregation/disaggregation techniques,
error analysis and calibration with the aid of
soil erosion rate measurements across
Europe. A database will be compiled on
existing soil erosion measurements from
plots and small catchments across Europe.

Distribution of
vegetation
resistance
Soil
erodibility
Sediment
transport as
soil erosion
Accumulation
over storms

Project phase 2 deals with validation and


comparison with other erosion risk
assessment methods across Europe and at
three different resolutions (field to
catchment, country and pan-European
scale). Linking existing data sets to model
parameters through transfer functions,
interpolation algorithms and statistical
methods will demonstrate the models
flexibility and robustness. The use of 10-daily
vegetation cover from NDVI and SPOT
Vegetation/HRVIR will enable calculated
estimates of seasonal variations in soil
erosion. Accurate spatial databases will be
compiled from existing information on
factors affecting erosion in Europe (climate,
soil, topography, land cover) and upgraded
using satellite imagery and computational
techniques.
Project phase 3 deals with application of the
model, development of a user-friendly
interface and establishment of user groups at
both national and European level.
Quantification of the erosion problem
enables evaluation of the possible effects of
future changes in climate and land use,
scenario analysis and impact assessment
according to cost-effectiveness, technical
feasibility, social acceptance and
implementability. End-user groups and
expert-user groups will actively participate in
model testing and in evaluating the projects
progress and results. Research networks will
be established in order to provide feedback
on the projects progress and results, and to
ensure continuation.

Annex III Background papers presented at the workshop

The integration of a plant growth model and erosion model into the Pesera model enables quantitative
forecasts of land cover and erosion

Plant growth model

Erosion model

Climate

Climate and distributions

Water balance

Runoff threshold

Veg./crop cover

f (soil, frost, moisture, cover)

Biomass production

Hillslope hydrology

SOM conversion

Sediment transport equation

To date, the model development has been


finalised (Figure A3.8). The model produces
a quantitative forecast of soil erosion and
plant growth, and therefore it has the
potential to respond explicitly and rationally
to changes in climate or land use, offering
great promise for scenario analysis and
impact assessment (Figure A3.9). Set against
this advantage, the model can only
incorporate the impact of past erosion where
this is measured and thus requires numerous
and good data sets needed for testing. The
model simplifies the set of processes
operating and may therefore not be
appropriate under particular local
circumstances. The Pesera model is currently
being calibrated and validated at different
resolutions and across different agroecological zones. Examples include
Andalucia and France (Kirkby and King,
1999; Kirkby et al., 2000). Test runs at high
resolution have demonstrated a satisfactory
goodness-of-fit (Gobin and Govers., 2001b).
Conclusions
In order to formulate a European soil
protection policy, policy-makers need to
know the area affected by soil erosion on a
regional scale, preferably on a pan-European
scale. In addition, the magnitude of soil
erosion in a particular area provides a useful
measure for formulating soil conservation
strategies.

Four methodologies have been reviewed that


present a regional assessment of soil erosion
by water. Both the Glasod maps and hot-spot
map were classified as methods based on
distributed point data, whereas the RIVM
and Corine maps were classified as factor- or
indicator-based maps. Methods based on
questionnaire surveys (Glasod) or erosion
measurement sites (hot spots) are
inadequate on their own. In addition,
differences between expert assessments and
measurement methods reduce the
comparability between the limited data
available. Methods based on factors or
indicators (RIVM, Corine) have the
immediate benefit of accessing distributed
climate data, soil maps, DEMs and land use
maps that are available on a European scale.
Corine makes explicit use of an adequate
range of relevant indicators for southern
Europe, but it is an imperfect
implementation, for historical reasons of
data availability, of a model (USLE) that is
now no longer considered as state of the art.
Although confined to southern Europe,
Corine gives the best indication of a regional
distribution of soil erosion of the four
methods reviewed.
The major problems with soil erosion are
firstly the temporal and spatial patchiness of
the phenomenon and secondly the
availability of widespread and long-continued
soil loss measurements or observations. In

91

Figure A3.9

92

Assessment and reporting on soil erosion

such cases, interpolations between available


sites are scientifically not justified. Current
regional assessment methods, such as Pesera
(Gobin et al., 1999), have therefore opted for
risk analysis combining plant growth, runoff
and sediment transport models. The Pesera
model produces quantitative forecasts of land
cover, runoff and soil erosion, and responds
explicitly and rationally to changes in climate
or land use. Since policy-makers have the
most direct impact on soil erosion through
land cover/use and land management
policies, a process modelling such as the
Pesera method enables further impact
assessment and scenario analysis.

Data availability for soil erosion


indicators at European level
Robert Jones
Commission of the European Communities Joint
Research Centre DG
Environment Institute, European Soil Bureau
Determining the causes of soil erosion
From the review of the current indicators for
soil selected by the EEA, it is concluded that,
from a scientific and technical standpoint,
the most appropriate indicator is the area
affected by erosion (see Chapter 2 in the
main text). However, because there is a
serious lack of direct measurements of soil
loss, by water and by wind, a surrogate
parameter or indicator is needed.
Conventional wisdom suggests that the area
actually affected by erosion should be directly
related to the area at risk from erosion
provided that the area at risk has been
determined using an appropriate model of
soil erosion together with the necessary
spatial data sets. Soil erosion takes place at
the field scale and the main problem is that
the digital data sets used to quantify the
factors causing erosion are usually too coarse
(in terms of spatial resolution) to enable
accurate estimation of soil losses at this scale.
Before considering modelling erosion risk
further, the concept of risk must be defined.
A risk is the chance of a bad consequence or
loss. Another definition of risk is the chance
that some undesirable event may occur. Risk
assessment involves the identification of the
risk, and the measurement of the exposure to
that risk.
The response to risk assessment may be to
initiate categorisation of the risk and/or to

introduce measures to manage the risk. In


some cases, the risk may simply be accepted.
In other cases, the priority will be to adopt a
mitigation strategy.
Such risk management, traditionally a
significant activity in the commercial sector,
e.g. the insurance industry, has now been
adopted in the environmental protection
arena. A review of soil erosion risk models
has been provided in Section 4.2 and in
Gobins and Kirkbys paper earlier in this
annex.
Modelling soil erosion
For developing suitable indicators for soil
erosion, a model-based approach is proposed
for assessing soil erosion risk. As stated above,
the availability of input data is a critical
selection criterion when assessing soil
erosion risk at the regional, national or
continental scale. Even though a wide variety
of models are available for assessing soil
erosion risk, most of them simply require so
much input data that applying them at these
scales becomes problematic.
The most widely used is probably the
universal soil loss equation (USLE)
developed by Wischmeier and Smith (1978).
The USLE has been used by Van der Knijff et
al. (1999, 2000) to estimate the risk of rill and
interrill erosion in Italy and Europe, using
the appropriate national and European data
sets.
A new harmonised European approach is
currently being developed under the EU fifth
framework programme. The project, panEuropean soil erosion risk assessment
(Pesera), will use the RDI model proposed by
Kirkby et al. (2000). It includes a module for
validation of the estimates at large scale.
Universal soil loss equation
The universal soil loss equation (USLE)
(Wischmeier and Smith, 1978) has used
widely because it is one of the least data
demanding erosion models that has been
developed and it has been applied at
different scales. The USLE is a simple
empirical model, based on regression
analyses of soil loss rates on erosion plots in
the USA. The model is designed to estimate
long-term annual erosion rates on
agricultural fields. Although the equation has
many shortcomings and limitations, it is
widely used because of its relative simplicity
and robustness (Desmet and Govers, 1996).
It also represents a standardised approach.

Annex III Background papers presented at the workshop

Soil erosion is estimated using the following


empirical equation:

A = R K L S C

(1)

Where:
A:
R:
K:
L:
S:
C:

Mean (annual) soil loss


Rainfall erosivity factor
Soil erodibility factor
Slope factor
Slope length factor
Cover management factor

The data sources that were used to estimate


the various USLE factors are summarised in
Figure 1.
Rainfall erosivity
The USLE rainfall erosivity factor (R) for any
given period is obtained by summing for
each rainstorm the product of total storm
energy (E) and the maximum 30-minute
intensity (I30). Unfortunately, these figures
are rarely available at standard
meteorological stations. Moreover, the
workload involved would be rather high for
any national or continental assessment.
Fortunately, long-term average R values are
often correlated with more readily available
rainfall figures like annual rainfall or the
modified Fourniers index (Arnoldus, 1978).
A similar approach was used to estimate R for
the whole of Europe (Van der Knijff et al.,
2000).
Soil erodibility
The K factor is defined as the rate of soil loss
per unit of R as measured on a unit plot
(Wischmeier plot). It accounts for the
influence of soil properties on soil loss

during storm events (Renard et al., 1997). In


the Corine study, soil texture, depth and
stoniness were used to estimate erodibility.
Working at the Member State and European
levels, Van der Knijff (1999, 2000) essentially
used soil texture. At the European level, Le
Bissonais and Daroussin (2001) have
developed a set of pedotransfer rules to
interpret the Soil Geographical Database of
Europe for soil erodibility and soil crusting.
Slope and slope length
The slope and slope length factors (S and L,
respectively) account for the effect of
topography on soil erosion. It can be
estimated from a digital elevation model
(DEM).
Vegetation cover and management
The cover and management C factor is
defined as the ratio of soil loss from land with
specific vegetation to the corresponding soil
loss from continuous fallow (Wischmeier and
Smith, 1978). Its value depends on vegetation
cover and management practices. Van der
Knijff et al. (1999, 2000) estimated C using a
combination of satellite imagery and a land
cover database based on Corine. Of those
factors used in the USLE, it is probably the
one accounting for the most variation in soil
loss.
Soil erosion risk assessments
A proper validation of results obtained from
applying an erosion model is hardly possible
on a small scale (e.g. national, continental).
Nevertheless, Van der Knijff et al. (1999,
2000) offer some comments on the general
pattern shown on the maps of erosion risk
that were produced for Italy and Europe
using the universal soil loss equation (USLE).

Flowchart for creating an USLE-based erosion risk map

MARS Meteo
Database

Annual
Rainfall

Classify
European Soil
Database

Surf.Texture +
Parent Mat.

93

A=RKLSC
Elevation
Model (1 km)

Slope

LS

NOAA-AVHRR +
CORINE

Cover Factor

Erosion
Risk
Map

Figure A3.10

94

Assessment and reporting on soil erosion

The most apparent contrast is between the


north and south in Europe. In general, soil
erosion risk seems to be underestimated for
most of northern Europe. This is mainly
caused by the rainfall erosivity factor, whose
predicted values are generally much lower
for northern Europe than for the south. Even
though rainfall in the north is less
aggressive compared to the south, the
differences shown on the map appear to be
too extreme. In southern Europe, the risk
assessments do not take past erosion into
consideration. For example, some areas
identified as being at high risk have already
been seriously eroded and the chances of
further soil loss are much reduced.
However, the work of Van der Knijff et al.
(2000) is important because it is one of the
first attempts to produce digitally a map of
soil erosion risk by rill and interrill erosion
for the whole of Europe. Its value lies in the
fact that the estimates of erosion risk are
based on standardised, harmonised data sets
for Europe. Moreover, the model output can
be estimates of actual soil erosion, by taking
crop/vegetation cover into account, and
estimates of potential erosion, by excluding
the cover factor. These results, together with
Figure A3.11
Source: Van der Knijff et
al., 2000.

Potential soil erosion risk in Europe

output from other modelling approaches,


should now provide a basis for defining a new
more physically based set of soil erosion
indicators for environmental auditing.
There is scope for major improvements in
such modelling by using more detailed
digital elevation data, better representation
of rainfall erosivity (i.e. more detailed rainfall
measurements), and satellite data that have
better spectral and geometric characteristics
than the NOAA-AVHRR data that are
currently available. Ideally, multi-temporal
satellite imagery should be used in order to
account for the interaction between
vegetation growth and senescence over the
year, and rainfall. Finally, more detailed soil
data are required (especially soil depth, stone
volume and surface texture).
Given such improvements in the basic data
sets, the output from a model such as the
USLE could provide a valuable indicator of
soil erosion. Some of the indicators currently
proposed such as fertiliser consumption,
farm structure and sediment transport could
then be useful but only for modifying a
standardised estimation of soil loss
determined using a standard model.

Annex III Background papers presented at the workshop

Environmental indicators for soil erosion


Environmental indicators for soil erosion are
proposed in Table 1. Examples of soil
erodibility and actual soil erosion risk

95

aggregated by catchment are provided in


Figures 3 and 4. Figure 2 shows the potential
soil erosion risk in Europe.

Soil erodibility class across Europe

Figure A3.12
Source: ESB, INRA.

Actual soil erosion risk aggregated by catchment

Figure A3.13
Source: ESB, INRA.

Yes

Yes

Yes

Vegetation/crop cover

Erosion control
measures

Yes

Yes

No

No

Yes

Yes

Yes

Soil storage capacity

Soil organic matter

Yes

In part

Topography (slope
length)

Yes

Yes

Soil crusting

Yes

Yes

In part

Yes

???

Erosivity

Yes

Yes

Soil drainage

Yes

Yes

Yes

Soil erodibility

Soil depth

Yes

Soil erosion risk

Soil texture

Yes

Easy to
interpret

Representative

Actual soil erosion

Utility

Policy
relevant

Yes

Yes

Yes

Yes

Yes

Yes

Yes (?)

Yes

Yes

Yes

Yes

Yes

???

Comparable

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Scientific/
Technically

Analytical
soundness

Rarely available

National/EU

National/EU

National/EU

National/regional

National/regional

National/regional

National/regional

National/regional

National/regional

National/regional

Nationally

Rarely available

Data available

Measurability

In part

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Yes

Yes

In part

Documented

No

Yes

Probably

Yes

Probably

Probably

Probably

Probably

Probably

Probably

Yes

Periodically

No

Updated

Direct

Simple

Simple

Simple

Simple

Simple

Simple

Simple

Simple

Simple

Complex

Complex

Direct

Effect

Usually piecemeal

Dichotomy between intensive indoor and outdoor


stocking

Updated by new data becoming available

Data for actual and estimated (CGMS) yields

Updated by new data becoming available

Updated by new data becoming available

Updated by new data becoming available

Updated by new data becoming available

Updated by new data becoming available

Updated by new data becoming available

Can be calculated from a standard model

Estimated from standard model e.g. USLE,

Extent not known, expensive to measure

Comments

Table A3.7

Environmental
Indicator

96
Assessment and reporting on soil erosion

Proposed environmental indicators and data availability at a European level

Annex IV Soil erosion glossary

Annex IV Soil erosion glossary


Term

Definition

Accelerated erosion

Erosion in excess of natural rates, usually as a result of anthropogenic activities

Actual erosion risk

The inherent risk of erosion under the current land use or vegetation cover

Actual erosion

Measured erosion

Source
1

Bank erosion

Erosion of the riverbank

Coastal erosion

Erosion of the coast, resulting in the retreat of the coastline

Erosion

(i) The wearing away of the land surface by rain or irrigation water, wind, ice or
other natural or anthropogenic agents that abrade, detach and remove
geologic parent material or soil from one point on the earths surface and
deposit it elsewhere, including such processes as gravitational creep and socalled tillage erosion; (ii) the detachment and movement of soil or rock by
water, wind, ice or gravity.

Erosion risk

Risk of erosion. It can be used as a surrogate indicator of actual erosion

Geological erosion

The normal or natural erosion caused by natural weathering or other


geological processes. Synonymous with natural erosion over a geologic time
frame or large geographic area

Gully erosion

The erosion process whereby water accumulates and often recurs in narrow
channels and, over short periods, removes the soil from this narrow area to
considerable depths, often defined for agricultural land in terms of channels
too deep to easily ameliorate with ordinary farm tillage equipment, typically
ranging from 0.5 m to as much as 25 to 30 m

Interrill erosion

The removal of a fairly uniform layer of soil on a multitude of relatively small


areas by splash due to raindrop impact and by sheet flow

Mass movement

Dislodgement and downslope transport of soil and rock material as a unit


under direct gravitational stress. The process includes slow displacements such
as creep and solifluction, and rapid movements such as landslides, rock slides,
and falls, earthflows, debris flows and avalanches. Agents of fluid transport
(water, ice, air) may play an important, if subordinate, role in the process

Potential erosion risk The inherent risk of erosion, irrespective of current land use or vegetation
cover
Rill erosion

An erosion process on sloping fields in which numerous and randomly


occurring small channels of only several centimetres in depth are formed; rills
can be obscured by tillage

Risk

The potential for realisation of unwanted, adverse consequences to human life,


health, property or the environment; estimation of risk is usually based on the
expected value of the conditional probability of the event occurring and the
consequence of the event given that it has occurred

Risk analysis

A detailed examination including risk assessment, risk evaluation and risk


management alternatives, performed to understand the nature of unwanted,
negative consequences to human life, health, property or the environment; an
analytical process to provide information regarding undesirable events; the
process of quantification of the probabilities and expected consequences for
identified risks

Risk assessment

The process of establishing information regarding acceptable levels of a risk


and/or levels of risk for an individual, group, society or the environment

Risk estimation

The scientific determination of the characteristics of risks, usually in as


quantitative a way as possible. These include the magnitude, spatial scale,
duration and intensity of adverse consequences and their associated
probabilities as well as a description of the cause and effect links

Saltation

A particular type of momentum-dependent transport involving: (i) the rolling,


bouncing or jumping action of soil particles 0.1 to 0.5 mm in diameter by wind,
usually at a height < 15 cm above the soil surface, for relatively short distances;
(ii) the rolling, bouncing or jumping action of mineral grains, gravel, stones or
soil aggregates effected by the energy of flowing water; (iii) the bouncing or
jumping movement of material downslope in response to gravity

Sediment

Transported and deposited particles or aggregates derived from rocks, soil or


biological material

Sedimentation

The process of sediment deposition

97

98

Assessment and reporting on soil erosion

Term

Definition

Sheet erosion

The removal of a relatively uniform thin layer of soil from the land surface by
rainfall and largely unchannelled surface runoff (sheet flow)

Soil erosion

Erosion of the soil. Soil erosion consists in the removal of soil material by water
or wind. It is a natural phenomenon but it can be accelerated by human
activities

Tillage erosion

The downslope displacement of soil through the action of tillage operations

Water erosion

The breakdown of solid rock into smaller particles and its removal by water. As
weathering, erosion is a natural geological process, but more rapid soil erosion
results from poor land use practices, leading to the loss of fertile topsoil and
to the silting of dams, lakes, rivers and harbours

Wind erosion

The breakdown of solid rock into smaller particles and its removal by wind. It
may occur on any soil whose surface is dry, unprotected by vegetation (to bind
it at root level and shelter the surface) and consists of light particles. The
mechanisms include straightforward picking up of dust and soil particles by the
airflow and the dislodging or abrasion of surface material by the impact of
particles already airborne

Sources
1

Glossary of soil science terms, Soil Science Society of America (SSSA), 1998

Glossary of risk analysis terms. Society for Risk Analysis


http://sra.org/glossary.htm

EEA glossary of environmental terms

Source

Annex V Processes of soil erosion

99

Annex V Processes of soil erosion


Over 90 % of non-glacial landscapes consist
of soil-covered hillslopes, with the remainder
being river channels and flood plains.
Although soil covered surfaces are not
generally the most active part of the
landscape, they provide almost all of the
material, which eventually leaves a river
catchment through the channel ways. The
processes by which material is weathered and
transported to the streams are therefore vital
to an understanding of how the landscape
transports weathered debris on hillslopes
(the regolith) and delivers sediment to
stream channels. Agriculture strongly affects
the rate and types of hillslope processes, and
the way in which farmland is managed can
dramatically influence whether soil erosion
remains at an acceptable level, or is increased
to a rate which leads to long-term and
perhaps irreversible degradation of the soil.
Slope sediment transport processes are of
two very broad types, first the weathering and
second the transport of the regolith. Within
each of these types, there are a number of
separate processes, which may be classified by
their particular mechanisms into groups
(Table 1), although many of these processes
occur in combination. Most slope processes
are greatly assisted by the presence of water,
which helps chemical reactions, makes
masses slide more easily, carries debris as it
flows and supports the growth of plants and

animals. For both weathering and transport,


the processes can conveniently be
distinguished as chemical, physical and
biological.
An additional important anthropogenic
process is tillage erosion, which is the result
of ploughing, either up- and downslope or
along the contour. Each time the soil is
turned over, there is a substantial movement
of soil. Up- and downhill ploughing produces
a direct downhill component of movement as
the turned soil settles back. Contour
ploughing can move material either up and
down, according to the direction in which
the plough turns the soil. Contour ploughing
in which the soil is turned downhill moves
approximately 1 000 times as much material
as soil creep. Contour ploughing in both
directions (soil turned uphill and then
downhill or vice-versa), or ploughing up- or
downhill produces a smaller net movement,
but the overall rate is still about 100 times
greater than natural soil creep. Sediment
transport is more rapid using modern heavy
machinery than with primitive ploughs, but it
is clear that tillage erosion may have been
responsible for more soil movement in the
last few centuries than natural soil creep
during the whole of the Holocene. The
accumulated effect is often seen in the buildup of soil behind old field boundaries.

Classification of the most important hillslope processes


Weathering processes

Transport processes

Chemical

Mineral weathering

Leaching
ionic diffusion

S
T

Physical

Freeze-thaw
Salt weathering
Thermal shattering

Mass movements
Landslides
Debris avalanches
Debris flows
Soil creep
Gelifluction
Tillage erosion
Particle movements
Rockfall
Through-wash
Rainsplash
Rainflow
Rillwash

S
S
S
T
T
T
S
T
T
T
T

Faunal digestion
Root growth
Microbial activity

Biological mixing (often included within soil creep)

Biological

Types: T = transport limited; S = supply limited removal (see below).

Type
(S/T)

Table A4.1

100

Assessment and reporting on soil erosion

Table A4.2

Types of soil erosion by water


Transportation

Mode

Detachment by

Through the air

In overland flow

Raindrop impact

Rainsplash

N/A

Overland flow traction

Rainflow

Rillwash gully erosion

Soil erosion by water


Although a small amount of material is
washed through the soil, the more important
erosion processes take place at the surface.
Material may be detached by two processes,
raindrop impact and flow traction; and
transported either by saltation through the
air or by overland water flow. Combinations
of these detachment and transport processes
give rise to the three main processes,
rainsplash, rainwash and rillwash, as
indicated in Table A4.2.
Raindrops detach material through the
impact of drops on the surface. For the
largest drops, the terminal velocity is 10 m s-1,
but they only attain this after falling through
the air for about 10 metres. If their fall is
interrupted by hitting the vegetation, drops
hit the ground at a much lower speed, and
have much less effect on impact. As drops hit
the surface, their impact creates a shock
wave, which dislodges grains of soil, or small
aggregates and projects them into the air in
all directions. The total rate of detachment
increases rapidly with rainfall intensity.
Where the raindrops fall into a layer of
surface water which is more than about 5 mm
thick, the impact of the drop on the soil
surface is largely lost.
Raindrop impact is also effective in breaking
down soil aggregates into constituent soil
particles. These particles are re-deposited
between aggregates on and close to the
surface, forming soil crusts, which seal the
surface, and limit infiltration by filling the
macropores between the aggregates. These
crusts may make the surface more resistant to
erosion, but their greatest importance is in
increasing runoff from storm rainfall.
Susceptibility to water erosion is closely
linked to the creation of soil crusts by rain
falling on unprotected surfaces, and the
destruction of crusts by tillage, freeze-thaw
and drying.
If water is flowing with sufficient force, it
exerts a force on the soil, which is sufficient
to overcome the resistance of soil particles.
Resistance is due to friction, which increases
with particle size, and cohesion between

grains, which increases with the specific


surface area of contact, and hence decreases
with particle size. Resistance is lowest for
small non-cohesive grains, particularly silts
and fine sands with a low clay content.
For rainsplash, grains are detached by drop
impact and jump through the air.
Transportation through the air, in a series of
hops, is able to move material both up- and
downslope, but there is a very strong
downslope bias on slopes of more than about
5 %. The net rate of downhill transportation,
therefore, increases with slope gradient, and
decreases with the grain size transported.
The rates of material transport by rainsplash
are generally low.
For rainflow, grains are detached by raindrop
impact, and carried farther than for
rainsplash within a thin layer of flowing
water. Both rainsplash and rainflow are most
significant in areas between small channels,
or rills, which form on the rapidly eroding
surface, and are commonly grouped together
as interrill erosion processes.
Where flow is sufficiently intense to entrain
soil particles directly, small channels or rills
are formed on the surface, and material is
eroded by rillflow, which is concentrated
along these drainage lines. In cultivated land,
resistance to erosion is commonly low within
the cultivated layer, but increases
considerably at the plough pan, which may
be a layer of increased resistance, and also
forms a transition to the undisturbed and
more consolidated unploughed soil beneath.
Rills therefore rarely penetrate beneath the
plough layer, and are generally obliterated by
later cultivation, as farmers seek to prevent
further erosion.
Under extreme storms, and where gradients
are at least locally steep, erosion may lead to
greater incision, forming gullies, which are
too large to be obliterated by normal tillage.
The development of gullies can fragment
farmland, and by steepening gradients to
adjacent fields, lead to rapid extension of a
gully network, which makes cultivation
impracticable. Remediation of gully systems

Annex V Processes of soil erosion

requires radical measures, including the


possible re-grading of entire landscapes.
Runoff is the most important direct driver of
severe soil erosion. Processes that influence
runoff must therefore play an important role
in any analysis of soil erosion intensity, and
measures that reduce runoff are critical to
effective soil conservation.
Perhaps the most important control on
runoff is the degree of crusting of the soil
surface, which has a very strong influence on
infiltration and therefore runoff rates. Of
secondary, but still major, importance is the
micro-topography of the soil surface and the
sub-surface soil structure, particularly the
presence or absence of macropores in the
form of cracks and/or voids between soil
aggregates. Micro-topography consists of
random roughness on the surface, together
with cultivation features such as plough
ridges and terracing. Both fine-scale microtopography and crusting evolve over the year,
in relation to tillage and the growth of crops
or uncultivated vegetation, their harvesting
or grazing, and the disposal of surface
residues.
Soil erosion is a natural process, occurring
over geological time, and most concerns

about erosion are related to accelerated


erosion, where the natural rate has been
significantly increased by human action.
These actions have generally been through
stripping of natural vegetation for
cultivation, indirect changes in land cover
through grazing and controlled burning or
wildfires, through re-grading of the land
surface and/or a change in the intensity of
land management, for example through
poor maintenance of terrace structures.
Increasing use of mechanised cultivation has
also led to a substantial increase in rates of
tillage erosion.
Erosion literature commonly identifies
tolerable rates of soil erosion, but these
rates usually exceed the rates that can be
balanced by weathering of new soil from
parent materials, and can only be considered
acceptable from an economic viewpoint.
Soil erosion by water is only one form of soil
degradation, which includes erosion by mass
movements, wind deflation/deposition and
other geomorphological processes. Soil is
also significantly degraded by salinisation,
particularly in arid areas, areas with salt-rich
parent materials and where water tables are
high.

101

102

Assessment and reporting on soil erosion

Annex V Processes of soil erosion

European Environment Agency


Assessment and reporting on soil erosion
2003 103 pp. 21 x 29.7 cm
ISBN 92-9167-519-9

103

You might also like